142
Mississippi State University Mississippi State University Scholars Junction Scholars Junction Theses and Dissertations Theses and Dissertations 8-7-2004 Examination of the Effects of Biosurfactant Concentration on Examination of the Effects of Biosurfactant Concentration on Natural Gas Hydrate Formation in Seafloor Porous Media Natural Gas Hydrate Formation in Seafloor Porous Media Charles E. Woods Follow this and additional works at: https://scholarsjunction.msstate.edu/td Recommended Citation Recommended Citation Woods, Charles E., "Examination of the Effects of Biosurfactant Concentration on Natural Gas Hydrate Formation in Seafloor Porous Media" (2004). Theses and Dissertations. 2097. https://scholarsjunction.msstate.edu/td/2097 This Graduate Thesis - Open Access is brought to you for free and open access by the Theses and Dissertations at Scholars Junction. It has been accepted for inclusion in Theses and Dissertations by an authorized administrator of Scholars Junction. For more information, please contact [email protected].

Examination of the Effects of Biosurfactant Concentration

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Examination of the Effects of Biosurfactant Concentration

Mississippi State University Mississippi State University

Scholars Junction Scholars Junction

Theses and Dissertations Theses and Dissertations

8-7-2004

Examination of the Effects of Biosurfactant Concentration on Examination of the Effects of Biosurfactant Concentration on

Natural Gas Hydrate Formation in Seafloor Porous Media Natural Gas Hydrate Formation in Seafloor Porous Media

Charles E. Woods

Follow this and additional works at: https://scholarsjunction.msstate.edu/td

Recommended Citation Recommended Citation Woods, Charles E., "Examination of the Effects of Biosurfactant Concentration on Natural Gas Hydrate Formation in Seafloor Porous Media" (2004). Theses and Dissertations. 2097. https://scholarsjunction.msstate.edu/td/2097

This Graduate Thesis - Open Access is brought to you for free and open access by the Theses and Dissertations at Scholars Junction. It has been accepted for inclusion in Theses and Dissertations by an authorized administrator of Scholars Junction. For more information, please contact [email protected].

Page 2: Examination of the Effects of Biosurfactant Concentration

EXAMINATION OF THE EFFECTS OF BIOSURFACTANT

CONCENTRATION ON NATURAL GAS HYDRATE

FORMATION IN SEAFLOOR POROUS MEDIA

By

Charles E. Woods, Jr.

A Thesis Submitted to the Faculty of Mississippi State University

in Partial Fulfillment of the Requirements for the Degree of Master of Science

in Chemical Engineering in the Dave C. Swalm School of Chemical Engineering

Mississippi State, Mississippi

August 2004

Page 3: Examination of the Effects of Biosurfactant Concentration

Copyright by

Charles E. Woods, Jr.

2004

Page 4: Examination of the Effects of Biosurfactant Concentration

EXAMINATION OF THE EFFECTS OF BIOSURFACTANT CONCENTRATION ON

NATURAL GAS HYDRATE FORMATION IN SEAFLOOR POROUS MEDIA

By

Charles E. Woods, Jr.

Approved:

_________________________________ ______________________________ Dr. Rudy E. Rogers Dr. Kirk H. Schulz Professor of Chemical Engineering Professor of Chemical Engineering (Director of Thesis) Earnest W. Davenport, Jr. Chair,

Director of Dave C. Swalm School of Chemical Engineering,

(Committee Member) _________________________________ ______________________________ Dr. Hossein Toghiani Dr. W. Todd French Associate Professor of Chemical Engineering Assistant Research Professor of (Committee Member) Chemical Engineering (Committee Member)

______________________________ Robert P. Taylor Interim Dean James Worth Bagley College College of Engineering

Page 5: Examination of the Effects of Biosurfactant Concentration

Name: Charles E. Woods, Jr.

Date of Degree: August 7, 2004

Institution: Mississippi State University

Major Field: Chemical Engineering

Major Professor: Dr. Rudy E. Rogers

Title of Study: EXAMINATION OF THE EFFECTS OF BIOSURFACTANT CONCENTRATION ON NATURAL GAS HYDRATE FORMATION IN SEAFLOOR POROUS MEDIA

Pages in Study: 128

Candidate for Degree of Master of Science of Chemical Engineering Various porous media were tested with biosurfactant solutions (rhamnolipid or

Emulsan) at concentrations ranging from 0 ppm to 1000 ppm. The biosurfactant

solutions in the presence of porous media often showed substantial gas hydrate

catalyzation, localization on selected surfaces, and/or specific hydrate form (massive,

stratified, dispersed.)

At 1000-ppm concentrations of rhamnolipid, all porous media surfaces exhibited

the same hydrate formation rate increase of 187% over the control. The curves generated

for rhamnolipid or Emulsan concentration versus peak hydrate formation rate resembled

certain classical adsorption curves.

Bentonite and aragonite showed hydrate catalyzation properties with or without

biosurfactants. The preference for hydrate formation on porous media surfaces (no

surfactant) was: Bentonite/nontronite > aragonite/stainless steel > Ottawa sand/kaolinite.

Porous media/biosurfactant concentration combinations play marked roles in the types of

Page 6: Examination of the Effects of Biosurfactant Concentration

gas hydrates formed: massive, dendritic, or needle-like. The research helps to explain the

vast occurrence of gas hydrates in ocean sediments.

Page 7: Examination of the Effects of Biosurfactant Concentration

-ii-

DEDICATION

I would like to dedicate this research and the culmination of this paper to my

parents, Charles and Barbara, my sister Amy, and especially to my understanding wife,

Marcia. Good things come to those who wait.

Page 8: Examination of the Effects of Biosurfactant Concentration

-iii-

ACKNOWLEDGMENTS

I would like to express my gratitude to all those who had a hand, directly or

indirectly, in the materialization of this study. I am indebted to Dr. Rudy E. Rogers for

his patience, dedication, and guidance through the many pitfalls of this project. I would

also like to thank my thesis committee of Dr. Kirk H. Schulz, Dr. Hossein Toghiani, and

Dr. W. Todd French for their time, effort, suggestions, and their inestimable help. I

would like to thank my coworkers, Jennifer Dearman, Ding Tao, and Dr. Gouchang

Zhang, for their rhetoric and support. Finally, I would like to show my deep appreciation

to all my friends and comrades, especially Transito Macias and Katrina Parker.

Page 9: Examination of the Effects of Biosurfactant Concentration

-iv-

TABLE OF CONTENTS

PAGE

DEDICATION.............................................................................................................. ii

ACKNOWLEDGMENTS ............................................................................................ iii

LIST OF TABLES........................................................................................................ vii

LIST OF FIGURES ...................................................................................................... viii

CHAPTER

I. INTRODUCTION................................................................................................. 1

II. LITERATURE REVIEW..................................................................................... 8

HYDRATE STRUCTURE ............................................................................................... 8 Structure I ............................................................................................................ 9 Structure II........................................................................................................... 9 Structure H .......................................................................................................... 10

GAS HYDRATE SOURCES ........................................................................................... 11 Thermogenic........................................................................................................ 12 Biogenic .............................................................................................................. 13

BIOSURFACTANTS...................................................................................................... 14 Rhamnolipid ........................................................................................................ 15 Emulsan............................................................................................................... 17

HYDRATE INHIBITION, PROMOTION, AND BIOSURFACTANTS..................................... 19 POROUS MEDIA.......................................................................................................... 20 POROUS MEDIA, SAND .............................................................................................. 21 POROUS MEDIA, CLAYS............................................................................................. 22

Kaolinite .............................................................................................................. 23 Bentonite ............................................................................................................. 24 Nontronite............................................................................................................ 27

POROUS MEDIA, MINERALS....................................................................................... 27

III. THEORETICAL BACKGROUND.................................................................... 29

SURFACTANTS ........................................................................................................... 29

Page 10: Examination of the Effects of Biosurfactant Concentration

-v-

CHAPTER PAGE

BIOSURFACTANTS...................................................................................................... 31 Rhamnolipid ........................................................................................................ 32 Emulsan............................................................................................................... 33

ADSORPTION THEORY................................................................................................ 34 Adsorption Classification.................................................................................... 34 Adsorption in Soils.............................................................................................. 37 Adsorption and Gas Hydrates ............................................................................. 38

INDUCTION TIME & NUCLEATION THEORY ............................................................... 39 GAS HYDRATE FORMATION RATE (KINETICS)........................................................... 41 PHASE EQUILIBRIA..................................................................................................... 43

Gas Gravity ......................................................................................................... 43 Distribution Coefficient Method ......................................................................... 44 Statistical Thermodynamics ................................................................................ 45

OVERALL MECHANISM .............................................................................................. 46

IV. EXPERIMENTAL METHODS ......................................................................... 50

HYDRATE EXPERIMENTAL SETUP .............................................................................. 50 HYDRATE PREPARATION............................................................................................ 54 ADSORPTION PREPARATION....................................................................................... 56 EQUIPMENT................................................................................................................ 58

Mass Balance....................................................................................................... 58 Constant Temperature Bath................................................................................. 59 Equilibration Bath ............................................................................................... 59 Sonicating Horn................................................................................................... 59 Surface Tensiometer............................................................................................ 60 pH meter .............................................................................................................. 60 Reaction Vessel ................................................................................................... 60 Sample Container ................................................................................................ 61 RTD probes ......................................................................................................... 61 Pressure Transducer ............................................................................................ 61 Linear Power Supply........................................................................................... 62 Pressure Relief Valve .......................................................................................... 62 Data Acquisition System..................................................................................... 62 Digital Camera .................................................................................................... 63

MATERIALS.............................................................................................................. 63 Rhamnolipid ........................................................................................................ 63 Emulsan............................................................................................................... 63 Ethanol ................................................................................................................ 63 Natural Gas.......................................................................................................... 64 Ottawa Sand ........................................................................................................ 64 Bentonite Clay..................................................................................................... 64 Kaolinite Clay ..................................................................................................... 64 Nontronite Clay ................................................................................................... 64

Page 11: Examination of the Effects of Biosurfactant Concentration

-vi-

CHAPTER PAGE

Aragonite ............................................................................................................. 65

V. RESULTS & DISCUSSION................................................................................ 66

SCOPE OF RESULTS .................................................................................................... 66 NATURAL GAS HYDRATE FORMATION RATE............................................................. 66 EFFECT OF BIOSURFACTANT CONCENTRATION ON FORMATION RATE....................... 70

Rhamnolipid Concentration Effects on Formation Rate ..................................... 71 Emulsan Concentration Effects on Formation Rate ............................................ 73

EFFECT OF POROUS MEDIA ON FORMATION RATE..................................................... 76 ADSORPTION OF BIOSURFACTANTS ON POROUS MEDIA............................................. 78 ADSORPTION AND BIOSURFACTANT CONCENTRATION RELATED TO FORMATION ..... 81 INDUCTION TIME........................................................................................................ 87 HEAT AND GAS TRANSFER EFFECTS ON FORMATION RATE ....................................... 90 EFFECT OF ELECTROLYTES ON FORMATION RATE ..................................................... 92 GAS HYDRATE IN POROUS MEDIA, PREFERENCE TRENDS ......................................... 93 GAS HYDRATE PACKAGING, BIOSURFACTANT ORDERING......................................... 96 DISPERSED SEDIMENT IN MASSIVE HYDRATES.......................................................... 99

VI. CONCLUSIONS ................................................................................................ 101

ADSORPTION.............................................................................................................. 101 FORMATION RATE..................................................................................................... 103 HYDRATE INDUCTION ............................................................................................... 104 STRUCTURE AND PREFERENCE.................................................................................. 104 SCIENTIFIC SIGNIFICANCE......................................................................................... 105 SUMMARY................................................................................................................. 106

REFERENCES ............................................................................................................. 107

APPENDIX

A. EXPERIMENTAL DATA................................................................................... 114

B. PENG-ROBINSON CALCULATIONS.............................................................. 127

Page 12: Examination of the Effects of Biosurfactant Concentration

-vii-

LIST OF TABLES

TABLE PAGE

2.1. Biosurfactant Classifications and Examples [26, p. 8] ..................................... 14

5.1. Experimental Matrix .......................................................................................... 67

5.2. Biosurfactant Selective Adsorption Test ........................................................... 79

5.3. Induction Time................................................................................................... 89

5.4. Effect of Electrolytes on Gas Hydrate Formation.............................................. 92

A.1. Experimental Plan ............................................................................................. 115

A.2. Surface Tension of Rhamnolipid at Room & Hydrate Temperature ............... 118

A.3. Surface Tension of Emulsan at Room Temperature ......................................... 118

A.4. Heat & Mass Transfer Effects on Hydrate Formation (Fig. 5.10) .................... 120

A.5. Effect of Rhamnolipid on Ottawa Sand, Averaged (Fig. 5.6)........................... 120

A.6. Effect of Rhamnolipid on Ottawa Sand/Bentonite, Averaged (Fig. 5.7).......... 121

A.7. Effect of Rhamnolipid on Ottawa Sand/Kaolinite, Averaged (Fig. 5.8)........... 121

A.8. Effect of Emulsan on Ottawa Sand, Averaged (Fig. 5.4) ................................. 121

A.9. Effect of Emulsan on Ottawa Sand/Bentonite, Averaged (Fig. 5.4)................. 122

A.10. Effect of Emulsan on Ottawa Sand/Kaolinite, Averaged (Fig. 5.4) ............... 122

A.11. Effect of Rhamnolipid on Varied Surfaces, Averaged (Fig. 5.3, 5.5) ............ 123

Page 13: Examination of the Effects of Biosurfactant Concentration

-viii-

LIST OF FIGURES FIGURE PAGE

2.1. Geometry of Gas Hydrates................................................................................. 11

2.2. Chemical Structure of Rhamnolipid .................................................................. 16

2.3. Emulsan Unit Structure [34] .............................................................................. 18

2.4. Structure of Sand................................................................................................ 21

2.5. Kaolinite Structure [47, p. 78] ........................................................................... 24

2.6. Sodium Montmorillonite Structure [47, p. 84] .................................................. 25

2.7. Aragonite Orthorhombic Structure [52]............................................................. 28

3.1. Adsorption Isotherms [69, p. B-278] ................................................................. 35

4.1. Diagram of Hydrate Sample Cup (Drawn to Scale) .......................................... 51

4.2. Hydrate Formation Cell Photograph.................................................................. 52

4.3. Diagram of Hydrate Reactor Vessel (Not to Scale)........................................... 53

4.4. Filled Sample Cup.............................................................................................. 55

5.1. Definition of Peak Formation Rate .................................................................... 69

5.2. Effect of Rhamnolipid Concentration on Gas Hydrate Formation Rate............ 71

5.3. Hydrate Formation Rate at 1000 ppm Rhamnolipid.......................................... 73

5.4. Effect of Emulsan Concentration on Gas Hydrate Formation Rate................... 74

5.5. Effect of Sediment on Peak Formation Rate in Distilled Water........................ 77

5.6. Rhamnolipid Concentration Vs Peak Formation Rate in Ottawa Sand ............. 82

5.7. Rhamnolipid Concentration Vs Peak Formation Rate in Bentonite Clay.......... 83

5.8. Rhamnolipid Concentration Vs Peak Formation Rate in Kaolinite Clay .......... 85

5.9. Induction Time in Gas Hydrate Growth ............................................................ 87

5.10. Effect of Heat and Mass Transfer Limitation .................................................. 91

5.11. Preference of Gas Hydrates to Stainless Steel Over Silica (OS) ..................... 94

Page 14: Examination of the Effects of Biosurfactant Concentration

-ix-

FIGURE PAGE

5.12. Preference of Gas Hydrates to Smectites......................................................... 95

5.13. Preference of Gas Hydrates to Aragonite with No Surfactant Present ............ 96

5.14. Gas Hydrate Packing Arrangements................................................................ 97

5.15. Nontronite Dispersed Within Hydrate Matrix ................................................. 99

A.1. CMC of Rhamnolipid at Room & Refrigerated Temperature .......................... 119

A.2. ST Vs Concentration of Emulsan at Room Temperature ................................. 119

Page 15: Examination of the Effects of Biosurfactant Concentration

-1-

CHAPTER I

INTRODUCTION

Gas hydrates are clathrate compounds consisting of gas molecules occluded

within an array of hydrogen-bonded water molecules. Unlike ice’s lattice structure, the

lattice of a gas hydrate crystal consists of a three-dimensional pattern characterized by a

preference for regular shapes such as pentagons and hexagons. These regular patterns

lessen the strain of the water bond angle of 104.5o most effectively. When introduced at

sufficiently high pressures and suitably low temperatures, gas molecules will, through

weak van der Waals forces, occlude into cavities of the hydrogen-bonded water network.

Gas hydrate structure has been likened to many naturally occurring patterns that

help in visualizing how the non-stoichiometric compound appears. Two such structures

are the buckminsterfullerene, or buckyball, and a common soccer ball. While

geometrically different, these patterns help to understand the construction of simple gas

hydrate molecules. Each shape, including gas hydrates, incorporates the use of regular

polygonal shapes to construct the surface of a hollow, cage-like structure. To date,

hydrates are known to appear as one of three common configurations or structures:

Structure I (sI), Structure II (sII), or Structure H (sH). These structures will be

considered in greater detail in the following chapter.

Page 16: Examination of the Effects of Biosurfactant Concentration

-2-

After 1810 when Sir Humphrey Davy discovered chlorine hydrates [1], gas

hydrates were considered for more than a century to be a laboratory curiosity with no

known natural occurrences. Gas hydrates did not draw much consideration until the

1930’s when the oil and gas industry started to detect icy deposits in many of their

pipelines, both above ground and sub-sea [2, p. 851]. These crystals were natural gas

hydrates that formed when moisture in the lines encountered cold temperatures of

surrounding environments or reduced temperature from Joule-Thompson expansion

cooling.

Gas hydrates have been discovered in abundance within sediment on the ocean

floor and in permafrost areas where temperatures and pressures are conducive. Gases

from these hydrates come from both geothermal and biological sources. As a matter of

fact, Brooks estimated that nearly half of the gas hydrates discovered in the Gulf of

Mexico was thermogenic while the other half was biogenic [3, p. 409]. Meanwhile,

Kvenvolden contends that gas hydrates from biological activity dominate in permafrost

areas [3, p. 409]. In addition, many microbial and subsea floor organisms such as tube

worms and mussels have been associated around or within gas hydrates [4, p. 5143; 5].

Within the last ten years, hydrates have acquired a significant amount of attention,

not only from the oil and gas industry, but also from such disciplines as chemistry,

chemical engineering, geology, oceanography, energy conservation, and environmental

engineering. Gas hydrates play a significant role in many natural phenomena such as

seafloor stability. Carbon dioxide sequestration and natural gas storage in gas hydrates

hold promise for significant societal benefits. Gas hydrates might also be a means to

recover natural gas resources, demineralize water, and predict global temperature

Page 17: Examination of the Effects of Biosurfactant Concentration

-3-

changes. The potential exploitation of some facets of gas hydrates or benefits from

hydrate prevention in other cases make gas hydrates a very interesting and productive

topic.

One area that may effectively be exploited in the near future is the recovery of

natural gas from seafloor and permafrost gas hydrates or “hydrate farming.”

Conservative estimates of natural gas present in seafloor hydrates reveal that

approximately twice as much methane may be sequestered in natural gas hydrates as all

of the methane-equivalent fossil fuels discovered to date [6]. As energy supplies are

subject to political perturbations, hydrates become more of a viable option for countries

feeling the energy crunch. Two such countries that are counting on hydrates to contribute

to their energy needs soon are Japan and India [7, p. 913; 8, p. 344] where exorbitant

amounts of hydrates are thought to exist in the countries’ territorial waters. If these

natural gas hydrate fields can be economically farmed, then natural gas could possibly

replace many environmentally harsh fossil fuels in use today.

Closely associated with hydrate farming is CO2 sequestration or disposal. In

addition to natural gas hydrates, carbon dioxide hydrates form at conditions present on

the ocean floor. Investigations have been made to determine the feasibility of depositing

industrial CO2 gas, a greenhouse agent that is being produced at an escalating rate, on the

ocean floor in the form of CO2 hydrates [9, p. 1067]. More importantly, if this process

could in some way be coupled economically with the harvesting of natural gas hydrates,

the progression would generate a viable natural resource fuel with no apparent negative

environmental effect such as net CO2 production.

Page 18: Examination of the Effects of Biosurfactant Concentration

-4-

While natural gas stored in seafloor hydrates may be a potentially viable source of

fossil fuel, this natural phenomenon also suggests a conventional natural gas storage

method. By fully utilizing the close packing of hydrocarbon gas within the cavities of

gas hydrates, a theoretical volume-to-volume ratio of 180 standard cubic feet of methane

to 1 cubic foot of host water may be achieved. Recently, many groups have investigated

applying this fact. Current methods of natural gas storage (depleted reservoirs, salt dome

caverns, compressed natural gas, LNG) are either expensive or require specific geological

formations that many parts of the U.S. and other countries simply do not have. Also,

liquefied natural gas (LNG) and compressed methane are potentially dangerous sources

of explosions if a storage tank were to rupture. Natural gas hydrate storage would

effectively eliminate many of these safety and convenience concerns.

Two contrasting methods of natural gas hydrate storage have been proposed

recently. Gudmundsson patented a process in 1996 in which gas hydrates could be

filtered from a cold slurry and not only stored, but also efficiently transported from site to

site [10]. However, Gudmundsson’s method relies primarily on cold slurry filtering and

a mechanical packing system which may be cost prohibitive.

As an alternative, Zhong and Rogers proposed a method by which hydrates could

be used for peak load storage of natural gas from a process requiring no agitating,

filtering, or packing by employing surfactants to lower the surface tension of the solution

[11]. The surfactant achieved uniform spatial packaging of the gas hydrates by

adsorption on metal surfaces at the gas-water interface. Their system also induced

hydrate formation rates at a 700-fold increase over quiescent systems. While

Page 19: Examination of the Effects of Biosurfactant Concentration

-5-

transportation of this system would be a difficulty, the system seems economically

practical for on-site storage of natural gas.

Additionally, in 2002 Mao, et al., showed evidence of a hydrogen gas form of

hydrates [12, p. 2247]. This new discovery may have inherent implications in on-board

storage of hydrogen for fuel cells and has also been linked to deep space H2-H20

agglomerations in planets or other celestial bodies. Similarly, gas hydrates have been

investigated as a budding storage possibility for natural gas powered vehicles [13, p.

209].

When gas hydrate crystals form and agglomerate into discrete particles, any

solutes in the solution are expunged to the interstitial water of the packed hydrate

particles. This fact suggests that gas hydrates could be used to desalinate or demineralize

water sources. However, as Barduhn, et al. question the economics of such a process and

point out the dependence on the hydrating agent and the difficulty of separating

interstitial water from the hydrates [14, p. 176].

While beneficial and profitable potentials exist for gas hydrate use, many

difficulties subsist. Much work was done in the early 1930s by the oil and gas industry to

deter the formation of natural gas hydrates in production pipelines [15, p. 66]. The most

common approach was to introduce a chemical inhibitor such as methanol into the lines

to depress the equilibrium temperature for natural gas hydrates. However, this

methodology hurts the profitability of oil and gas production in deep waters offshore.

The oil and gas industry typically spends millions of dollars annually to prevent gas

hydrates from forming and plugging pipelines offshore.

Page 20: Examination of the Effects of Biosurfactant Concentration

-6-

To combat the problem, alternatives to thermodynamic inhibitors have been

proposed. In 2001, Huo, et al., investigated the effectiveness of using anti-agglomerates,

compounds that prevent hydrates from forming critical nucleation clusters, to prevent

pipeline plugging [16, p. 4980].

Another major problem is the effect that gas hydrates have on seafloor stability.

While gas hydrates are easily produced and are abundant in nature when the environment

is conducive, they are innately meta-stable compounds. Formation and decomposition of

seafloor hydrates may occur through shifts in temperature due to ocean currents, shifts in

geothermal gradients, or temperature changes triggered by deep-ocean drilling. Whatever

the mechanism may be, rapid decomposition of hydrates can cause enormous sub-sea

landslides on continental slopes or margins and can also trigger massive localized

releases of gas to the sea surface. Rapid dissociation of gas hydrates may also efface the

cement holding together many ocean floor sediments, turning the sediments into a low-

strength, non-cohesive mud. This type of event has serious implications to oil drilling

rigs or pipelines that may be anchored in these cemented sediments [17, p. 1791].

An increasing number of scientists have blamed a massive decomposition of

natural gas hydrates for the ending of several glacial ages and the extinction of many

species, including the disappearance of the dinosaurs and the conclusion of the last ice

age [18, p. 392; 19, pp. 357 - 358; 20, pp. 691 – 693; 21, p. 443]. In addition, Yevi and

Rogers have also referenced gas hydrates as a prospective cause for the disaster at Lake

Nyos in Africa in 1986 [13].

While irrefutable evidence is evasive at this time, there are significant indications

that gas hydrates have been in the past and may be in the future major contributors to

Page 21: Examination of the Effects of Biosurfactant Concentration

-7-

global climate change. Even if highly conservative estimates of the quantity of gas

hydrates are true, hydrates could potentially be the source of more greenhouse gas

(methane) than industry would produce (CO2).

Seafloor instabilities due to hydrates has prompted a massive research effort to

better understand the environment in which these compounds reside. Investigations into

hydrate interactions with sand, clays, and silt are ongoing. However, to a large extent,

mechanisms and conditions of hydrate formation on the ocean floor are still elusive.

Many questions remain as to why and where hydrates form in the depths of the

ocean. Acoustic devices and core sampling have increased knowledge substantially on

how to locate gas hydrates, but why gas hydrates form where they do and what influences

that formation is still an enigma. It has been demonstrated that certain porous media

facilitate the formation of gas hydrates, and it has also been shown that the effects of

surfactants on gas hydrate formation are also catalytic. The goal of this research is to

determine why hydrates form preferentially on surfaces and why surfactant solutions,

particularly biosurfactant solutions, seem to dictate this preference in certain cases.

Subsequently, a mechanism for the formation of hydrates in deep-sea sediment will be

proposed in the context of surfactant interaction or absence.

Page 22: Examination of the Effects of Biosurfactant Concentration

-8-

CHAPTER II

LITERATURE REVIEW

Hydrate Structure

Gas hydrates are a three-dimensional lattice of water, hydrogen-bonded into

regular polyhedra and stabilized by an occluded gas guest molecule that is encaged

through weak van der Waals forces. Water molecules form the vertices of the polyhedra

while line segments connecting two vertices represent hydrogen bonds. The gas hydrate

lattice has three known polymorphs designated as Structure I (sI), Structure II (sII), and

the more recently discovered Structure H (sH). The three structures differ in cubic

structure, cage diameter, and theoretical hydrate number (a ratio of water molecules

comprising the structure to the cages formed by the structure.)

Gas hydrates may be compared to ice Ih, a system of water molecules hydrogen

bonded to one another in a tetrahedral pattern forming puckered rings instead of discreet

planar sheets of ice [3, p. 25 – 27]. Unique to gas hydrates as opposed to ice, however, is

the occurrence of completely convex polyhedra that form hydrogen-bonded cages. While

ice under the same configuration would collapse, guest molecules within these cages or

within neighboring cages help establish structure rigidity and stability.

Page 23: Examination of the Effects of Biosurfactant Concentration

-9-

Structure I

Structure I hydrates form repetitive structures of pentagonal dodecahedra denoted

as 512 and tetrakaidecahedra denoted as 51262. This nomenclature follows the pattern

proposed by Jeffrey, imin , where mi is the number of faces of a single polyhedron and ni is

the number of edges in that face [3, p. 32]. The 12-hedron structure is the smaller of the

two geometries, is nearly spherical, and has a cage radius of 3.91Å [3, pp. 33 – 35]. The

smaller dodecahedra are comprised of 20 water molecules surrounding each guest

molecule while the larger tetrakaidecahedra are comprised of 24 water molecules

surrounding each guest molecule. The tetrakaidecahedron has an oblate structure

stemming from the orienting of two hexagons on either pole of the 14-hedron. The

average cavity radius for a tetrakaidecahedron is 4.33Å, allowing it to occlude ethane,

propane, and other non-spherically-shaped sI hydrate formers [3, pp. 36 – 37]. The sI

hydrate configuration consists of 8 guest molecules (6 large and 2 small) and 46 water

molecules arranged in a body-centered cubic unit cell. The hydrate number for Structure

I hydrates (theoretical host-to-guest ratio) is 5.75.

Structure II

Structure II gas hydrates assemble by face sharing of the pentagonal dodecahedra.

It is this face sharing that allows for the second cavity size, the hexakaidecahedron, not

present in sI or sH type hydrates. The tetrakaidecahedron is absent in sI or sH types. The

hexakaidecahedron is a 16-hedra comprised of 28 water molecules arranged into 12

pentagons and 4 hexagons (51264). The 16-hedron creates an almost spherical cage with a

radius of 4.68Å capable of supporting molecules as large as iso-butane and n-butane with

Page 24: Examination of the Effects of Biosurfactant Concentration

-10-

help from a smaller, stabilizing hydrate former such as methane [3, pp. 37 – 38].

Structure II hydrates fashion a diamond-centered cubic unit cell constructed of 24 guest

molecules (8 large and 16 small) and 136 hydrogen-bonded water molecules. Structure II

hydrates have a hydrate number of 5.67.

Structure H

In 1987, Ripmeester and Ratcliffe published work on a newly discovered structure

for gas hydrates [22, p. 8773 – 8776]. The new structure, designated Structure H or sH,

was known to occlude larger gas molecules than originally thought possible. This fact

occurs because larger, oblong 20-hedra cages are created within its structure. The 20-

hedron structure consists of twelve pentagonal faces along with eight hexagonal faces

(51268). Along with the 20-hedra and the pentagonal dodecahedra, there is a third cavity

size in each unit cell. Also unique to the sH type hydrate is another twelve-faced

polyhedron structure comprised of six pentagonal faces, three hexagonal faces, and three

bond-strained square faces, (435663). Due to the strain placed on the unit structure by the

435663 polyhedra, sH hydrates only occur if there is sufficient filling of the 20-hedra and

pentagonal dodecahedra to stabilize, or meta-stabilize, the clathrate. This fact has

prompted some to refer to sH type hydrate as a double hydrate, requiring a large and a

small guest molecule to form.

The sH hydrate structure is a hexagonal unit cell containing three pentagonal

dodecahedra, two of the other dodecahedra (435663), and one 20-hedron. The 435663

dodecahedra have a radius of 4.06Å while the larger 20-hedron has a radius of 5.71Å

allowing it to accommodate some cyclical and branched compounds. The sH type

Page 25: Examination of the Effects of Biosurfactant Concentration

-11-

hydrate has 34 water molecules per 6 guest molecules (theoretical) in each unit cell

giving a hydrate number of 5.67. A representation of each type of polyhedral cage

comprising gas hydrates is shown in Figure 2.1.

512 51262 51264 435663 51268

3.91Å 4.33Å 4.68Å 4.06Å 5.71Å

Figure 2.1. Geometry of Gas Hydrates Although each of the hydrate numbers, n, given above is below six, these values

are theoretical and rarely the case in nature. Gas hydrates can exist with only partial

filling of the hydrate cages, making them non-stoichiometric hydrates as opposed to

stoichiometric hydrate salts. Under natural conditions, however, the hydrate number may

be much larger than the theoretical value due to incomplete filling of the hydrate cages

[23, p. 4].

Gas Hydrate Sources

As stated previously, gas hydrate formation requires a concurrence of several

factors including sufficient water, sufficient gas concentration, proper temperatures, and

proper pressures. In nature, hydrates form prominently on the ocean floor and permafrost

areas where not only are temperatures and pressures adequate but water is in abundance.

Page 26: Examination of the Effects of Biosurfactant Concentration

-12-

Often the limiting factor is the proper gas concentration. Ocean floor and permafrost

natural gas hydrates can evolve from two major sources: thermogenic and biogenic.

Thermogenic

Thermogenic natural gas hydrates are abundant near ocean floor and permafrost

seeps. These gases are formed by catagenesis in the earth’s crust through high

temperatures and extremely high pressures, producing natural gas usually rich in ethane,

propane, and higher hydrocarbons. When fissures form in the ocean floor or near the

surface in the arctic permafrost, the gas bubbles or seeps through a large reservoir of

water where the gas may be enclathrated in gas hydrates.

Gas hydrates from thermogenic sources have both a characteristically high

amount of carbon-12 isotopes and a low ratio of methane to ethane and propane. The

Peedee Belemnite (PDB) standard is often used to distinguish between thermogenic gases

and biogenic gases and stems from the following relative reference.

( )( )

3

1213

1213

13 101 ×

−≡

PDB

Sample

CC

CC

Cδ (2.1)

Here 13C and 12C denote the different isotopes of carbon. The measurement is reported in

parts per thousand and is usually between –25% and –60% for thermogenic sources.

Equation 2.2 defines the methane-to-ethane and propane ratio.

( )32

1CC

CRMtEP +≡ (2.2)

Thermogenic gas sources typically have a ratio less than 100 [3, pp. 405 – 407].

Page 27: Examination of the Effects of Biosurfactant Concentration

-13-

Biogenic

Bacteria in ocean sediments or permafrost by contrast produce natural gas that is

predominantly methane. These microorganisms produce both methane and carbon

dioxide through a diagenesis process. In anoxic sediments, microorganisms are required

to reduce minerals within the sediment to sustain life. As the chemical composition of

the sediment changes with depth, the layers of microorganisms adapt to these

environments and reduce what minerals they have at their disposal. Oxidation and nitrate

reduction occur in the upper levels of sediment where dissolved oxygen and nitrate

compounds are abundant. As depth increases, sulfate reduction dominates in the region

where sulfate compounds are prominent. Finally, carbonate reduction and fermentation

becomes the method of choice as carbonate compounds increase in concentration. It is

from this region where biogenic gases are emitted into the upper layers of the sediment

where they are contacted with ample amounts of water for hydrate formation.

As a result of this diagenesis, gas sources produced by microorganisms are rich in

methane, giving RMtEP values (Eq. 2.2) of greater than 1,000. Biogenic sources also have

a considerably lower δ13C range of –55% to –85% [3, pp. 405 – 406]. This lower range

is indicative of microbes’ ability to metabolize the lighter 12C isotope rather than the 13C

isotope. As a result, methane produced by biogenic sources is high in the 12C isotope.

Biogenic sources are thought to be the dominant mode of generation of natural gas

hydrates in permafrost regions [3, p. 409].

Page 28: Examination of the Effects of Biosurfactant Concentration

-14-

Biosurfactants

Biosurfactants are naturally produced, surface-active compounds typically having

both hydrophilic and hydrophobic segments. Detergents are common examples of man-

made surfactants that function similarly to biosurfactants. Biosurfactants are natural

emulsifiers, solubilizers, surface tension reducers, and antimicrobial agents [24, p. 1009;

25, p. 1737]. They serve such purposes as bringing an immiscible carbon source to the

water-borne bacterial cell by solubilizing a mineral or organic compound.

Biosurfactants, acting as antibiotics, may also aid the parent microbe in competitive

exclusion where the producer is unaffected but the competition is exterminated. The

types of hydrophilic and hydrophobic groups a biosurfactant contains determines greatly

the activity and specificity of the biosurfactant. A list of the five common classifications

of biosurfactants, an example of each classification, and the accompanying parent

microbes are given in Table 2.1.

Table 2.1. Biosurfactant Classifications and Examples [26, p. 8] Classification Biosurfactant Example Parent Microbe

Hydroxylated and cross-linked fatty acids

DL-A-hydroxystearic acid Cornybacterium lepus

Polysaccharide-lipid complexes Emulsan Acinetobacter calcoaceticus

Glycolipids Rhamnolipid Pseudomonas aeruginosa

Lipoprotein-lipopeptides Surfactin Bacillus subtilis

Phospholipids DMPC Cornyebacterium insidiosum

Page 29: Examination of the Effects of Biosurfactant Concentration

-15-

Surfactants and biosurfactants have been recognized as good biological degraders

[24, p. 1009; 27, p. 3901]. Rhamnolipid in particular is known for its adeptness at

removing oil from sand, and Emulsan played an intricate part in the clean up of the

Exxon Valdez oil spill [28, p. 22; 29]. Surfactants, however, have only recently been

considered for use with gas hydrates. Zhong and Rogers have investigated the effect that

sodium dodecyl sulfate (SDS) has on gas hydrate formation [30]. Both Lee and

Kothapalli have recently studied the effect that biosurfactants may have on the formation

of natural gas hydrates in ocean sediments [26, pp. 78 – 81; 31, pp. 75 – 76].

Rhamnolipid

Rhamnolipid is a biosurfactant produced by the bacterial strain Pseudomonas

aeruginosa. Rhamnolipid can be mono-headed (consisting of a six-carbon sugar) or

multi-headed as well as mono-tailed (consisting of a seven-carbon alkyl) or multi-tailed.

The chemical structures of two types of rhamnolipid, one dual-headed and dual-tailed, the

other mono-headed and dual-tailed, are shown in Figure 2.2.

Page 30: Examination of the Effects of Biosurfactant Concentration

-16-

o

oH

CH

(CH2)6

CH3

CH2

C

o

o

o

oH

oH

oH

oH

o

o o

oHCH

(CH2)6

CH3

CH3

CH3

CH2

C

o

CH

(CH2)6

CH3

CH2

C

oHCH

(CH2)6

CH3

CH2

C

o o

o

oH

o

oH

oH

CH3

Figure 2.2. Chemical Structure of Rhamnolipid

Rhamnolipid is an anionic glycolipid that is highly surface-active as indicated by

its low critical micellar concentration (CMC) value of 18 parts per million (ppm) in water

at atmospheric conditions as a rhamnolipid mixture, or 60 ppm at atmospheric conditions

for purified mono-headed, dual-tailed rhamnolipid in water [32, p. 1995]. Rhamnolipid is

a micellar agent that orients with its hydrophilic saccharine head solvated by water and

with its hydrophobic lipid tail directed away from water, usually in a micelle (except at

acidic pH.)

Page 31: Examination of the Effects of Biosurfactant Concentration

-17-

Rhamnolipid’s morphology depends heavily on such factors as the pH of solution.

Since rhamnolipids are essentially carboxylic acids, they have more negatively charged

heads at higher pH values. This fact creates a morphological change from lamellae

(bilayer sheets) to vesicles (hollow bilayered, fluid-filled spheres) to micelles as pH

increases from 5.0 to 8.0 [33, p. 570 – 572].

Emulsan

Emulsan is an anionic biological emulsifier produced by the bacterial strain

Acinetobacter calcoaceticus. Emulsan is technically referred to as a biopolymer because

of its saccharine backbone that coils and twists into a tertiary structure that can emulsify

oil droplets. From this polysaccharide backbone, Emulsan has numerous lipid side-

chains giving it the classification of a polysaccharide-lipid complex. The U.S. Army has

heavily researched Emulsan in search of a biodegradable, environmentally friendly

detergent and degreaser [34]. However, due to its extraordinary size (980,000 molecular

weight) [35, p. 132] and folding abilities, Emulsan is not a micelle-forming surfactant.

The basic unit structure of the Emulsan polymer is shown in Figure 2.3.

Page 32: Examination of the Effects of Biosurfactant Concentration

-18-

oH COO

CH2OCO

AcNHHO o

n

oo

AcNH

o

o

o

oH AcNH

CH2OCO

Figure 2.3. Emulsan Unit Structure [34] In the above diagram, AcNH refers to a secondary amine containing an acetyl

group (COCH3) as well as being bound to the six-carbon, puckered sugar ring. The

jagged lines refer to the saturated and unsaturated lipid tails while the n denotes that the

structure is a repeating polymer chain of n subunits.

While emulsifying and degreasing properties are well known, the capabilities of

Emulsan to solubilize light n-alkane gases have not been reported. Kothapalli has

considered Emulsan for possibilities of hydrate catalysis in certain porous media [26, p.

68 – 73].

Page 33: Examination of the Effects of Biosurfactant Concentration

-19-

Hydrate Inhibition, Promotion, and Biosurfactants

Since the oil and gas industry first experienced gas hydrate blockage of pipelines,

the industry has been investigating ways to combat the perceived nuisance. Early on, two

methods to retard hydrate formation were employed, thermodynamic inhibition and

kinetic inhibition. Thermodynamic inhibition primarily shifted the equilibrium line for

gas-liquid water-hydrate stability while kinetic inhibition controlled the rate at which

hydrates formed, preventing detrimental plug formation in a given time frame. The

conventional method of thermodynamic inhibition, studied in depth by Hammerschmidt,

was to inject methanol which reduces the chemical potential of water [15, p. 66]. Some

thermodynamic inhibition occurs from electrolyte solutions such as saltwater solutions

[36, pp. 70 – 73]. Tohidi, et al., looked at predictive models of injecting methanol or

electrolytes to prevent pipeline blockage by hydrates. Recently Jager, et al., have shown

that the inhibitory effects of methanol and electrolytes together are greater than the sum

of their parts [37, pp. 34 – 37; 38, p. 27].

Lately a third method of hydrate inhibition, anti-agglomeration, has been

introduced. Anti-agglomerates do not affect the rate of hydrate crystallization as kinetic

inhibitors do; rather they hinder the hydrate crystals from coalescing into critical clusters

of hydrates and thus prevent precipitation. These chemicals, typically surfactants, have

an affinity for both the hydrate crystal and a liquid oil phase. Huo, et al., have

investigated many non-ionic commercial and synthesized surfactants for their anti-

agglomeration capabilities [16, pp. 4982 – 4983].

Of late, there has been a renewed interest in surfactants as a means of hydrate

promotion. Surfactants have been used to promote hydrates for natural gas storage

Page 34: Examination of the Effects of Biosurfactant Concentration

-20-

possibly as an alternative to LNG or compressed natural gas storage. For example,

sodium dodecyl sulfate has exhibited the ability to increase hydrate formation rate in a

non-stirred system by more than 700 fold, assist in agglomeration into a tightly packed

structure, and uptake more than 97% of the theoretical maximum uptake of methane into

hydrates [30, p. 4177; 39, pp. 5 – 9]. In addition, biosurfactants such as surfactin,

rhamnolipid, and Emulsan have been shown to promote or inhibit hydrate formation

under certain interactions with porous media [26, pp. 68 – 73; 31, pp. 75 – 76].

Porous Media

Many researchers have hypothesized about the effect that various porous media

may have on the formation, stability, and induction of gas hydrates [8, pp. 344 – 348; 26,

pp. 78 – 81; 40, pp. 237 – 239; 41, pp. 6492 – 6494; 42, pp. 977 – 980; 43, p. 3659; 44].

Recently, a connection between gas hydrate emergence and surfactants and biosurfactants

has been suggested [16, p. 4990; 26, pp. 78 – 81; 30, p. 4175; 31, pp. 98 – 100; 42, p.

973].

Kothapalli has made a connection between an increase in hydrate activity with the

presence of both porous media and biosurfactants [26, pp. 78 – 81]. Lanoil, et al., have

made the discovery of Proteobacteria, such as Psuedomonas aeruginosa, and

Actinobacteria within core samples taken from Gulf of Mexico gas hydrate mounds,

suggesting the presence of the precursors to biosurfactants used in the research for this

thesis [4, pp. 5146 – 5148].

Kothapalli also showed that natural gas hydrates have a preference to specific clay

surfaces in the presence of certain biosurfactants [26, pp. 68 – 73], and it is discussed in

Page 35: Examination of the Effects of Biosurfactant Concentration

-21-

following sections the preference of natural gas hydrates to the smectite clays or metal

surfaces over other surfaces.

Porous Media, Sand

Sand is comprised of a tetrahedral arrangement of silicon and oxygen in the

stoichiometric ratio 1:2 with a chemical structure SiO2. As such, the surfaces of sand

particles are negatively charged due to the dominance of oxygen’s two lone pairs of

electrons. A diagram of the tetrahedral structure of SiO2 is shown in Figure 2.4.

Sio

o

o oSi

o

o

oSi

o

oSio

o

oSi

oo

Sio

Si

o

o oSi

o

o

oSi

o

oSi

o

oSi

oo

SiSio

o

o oSi

o

o

oSi

o

oSio

o

oSi

oo

Sio

Si

o

oSi

o

o

Si

o

oSi

o

Si

oo

Si

Figure 2.4. Structure of Sand Sand’s tetrahedral pattern suggests a possibility of hydrogen bonding when in the

presence of fluids capable of such bonds. Literature suggests that while protons of water

are attracted to the surface of sand through hydrogen bonds, the bonding only occurs very

near the surface and breaks down beyond approximately three molecular layers [45, pp.

103 – 105].

Ottawa sand is a purified form of SiO2 with a narrow particle size distribution

making it amenable to research. Ottawa sand has a grain size of between 20 and 30 mesh

Page 36: Examination of the Effects of Biosurfactant Concentration

-22-

units. When compared to other minerals, sand has an unusually high porosity because of

its relatively large particle size, and it is noteworthy that Ginsburg, et al., state a

preference of gas hydrates for highly porous media [40, p. 237]. High-porosity, packed

media give gas hydrates more room to expand upon solidifying much like ice would.

The pH at which a metal oxide surface has a net zero charge is typically referred

to as the zero point of charge, pHZPC. At any pH value above the pHZPC, the surface will

have a net negative charge. Conversely, the surface will attain a net positive charge at

any pH value below the pHZPC. Sand has a pHZPC range of 2.9 – 3.0 meaning that, at

moderate conditions (between pH 6.0 and 8.0), sand will acquire a net negative charge

[46, p. 93]. While water will loosely solvate the surface of a negatively charged sand

particle, an anionic or poly-anionic biosurfactant will not associate with the sand surface

under these conditions. As a result, for moderate pH, no appreciable adsorbed micelles

from anionic surfactants should exist on the surface of sand.

Porous Media, Clays

Clays are typically dually defined by their size and/or by their structures. While

many technical sources refer to clay minerals as layer silicates or phyllosilicates, more

general definitions prescribe that clays consist of any inorganic material less than 2 µm in

diameter [47, p. 76]. To avoid ambiguity, clays or clay minerals will be defined

exclusively as layered silicates within this paper. Typical nomenclature refers to layered

silicates by the ratio of their tetrahedral sheets to their octahedral sheets. Hence,

pyrophyllite, a clay consisting of one aluminum octahedral sheet sandwiched between

two silicon tetrahedral sheets, is cited as a 2:1 layer silicate.

Page 37: Examination of the Effects of Biosurfactant Concentration

-23-

Agricultural soil typically consists of six clay groups with some exceptions:

kaolins, smectites, vermiculites, micas, chlorites, and soil clays [47, p. 83 – 88]. Only

two of these groups will be examined at length here, the kaolin and smectite groups. Of

these two groups, the kaolinite mineral will be examined from the kaolin group, and

montmorillonite and nontronite will be investigated from the smectite group.

Kaolinite

Kaolinite is a white powder commonly referred to as china clay due to its role in

creating fine china. Kaolinite is a 1:1 layered silicate with alternating silicon oxide

tetrahedral sheets and aluminum hydroxide octahedral sheets. These two layers connect

at their basal planes through hydrogen bonding. A diagram of kaolinite structure is

shown in Figure 2.5. Kaolinite in its natural state is an extremely soft mineral with a

Mohs’ hardness of approximately 1. This fact is likely attributable to its preference to

form flakes and scales upon crystallization in the monoclinic system [48, p. 43].

Kaolinite has a specific weight of 2.6. Kaolinite has an overall general structure of

Al2Si2O5(OH)4 with spacing between its basal planes of 7.2Å. Kaolinite also has a pHZPC

value of 4.7 – 5.1, resulting in an overall negative surface at moderate pH values [46, p.

93].

Page 38: Examination of the Effects of Biosurfactant Concentration

-24-

Si

O

Al

OH

Hydrogen Bonding Region

Figure 2.5. Kaolinite Structure [47, p. 78]

Kothapalli investigated the effect on hydrate formation of kaolinite in the

presence and absence of assorted biosurfactants [26, pp. 68 – 73]. It was shown kaolinite

wetted with 1000-ppm biosurfactant solutions, including rhamnolipid and Emulsan,

decreased nucleation time, or induction time, in all cases. The rate of formation due to

biosurfactant interaction with kaolinite increased in all instances, most notably in the

cases of rhamnolipid and Emulsan at 1000 ppm. Also, rhamnolipid and Emulsan were

shown to adsorb on the surface of kaolinite to an unquantified extent as seen by increases

in surface tension due to kaolinite contact with biosurfactant solution.

Bentonite

Bentonite clay, named after its origin of discovery in Fort Benton, Wyoming, is a

member of the smectite group. The majority of bentonite is composed of

montmorillonite clay, a 2:1 layered silicate. The mainly aluminum hydroxide octahedral

sheet in montmorillonite is sandwiched between two silicon dioxide tetrahedral sheets.

Page 39: Examination of the Effects of Biosurfactant Concentration

-25-

The negative charge attained by the surface of the silicon dioxide basal plane is stabilized

by exchangeable cations, most notably sodium, calcium, and magnesium [47, pp. 83 –

84]. Other metals such as iron or magnesium commonly substitute aluminum within the

octahedral sheet in montmorillonite. These substitutions cause impurities in the basic

structure. Figure 2.6 shows the structural arrangement of a sodium-exchanged

montmorillonite platelet.

+ + +

+

Si

O

Al

OH

Na+

Figure 2.6. Sodium Montmorillonite Structure [47, p. 84]

Montmorillonite crystallizes as a grayish white powder in the rhombic system.

Like kaolinite, bentonite is a very soft mineral and may attain a waxy feel [48, pp. 43 –

44]. Also like kaolinite, montmorillonite is referred to as a “plastic” clay, taking the

shape of its container when wetted [47, p. 84]. Bentonite and montmorillonite are

notorious for their water adsorptive, or swelling, properties by which they can adsorb

several times their weight in water. Montmorillonite basal plane spacing may range from

Page 40: Examination of the Effects of Biosurfactant Concentration

-26-

9.6Å to more than 18Å but may reach distances of tens or hundreds of Ångstroms upon

swelling. Montmorillonite, also due to its swelling properties, has many industrial uses.

The clay is typically used as cat litter because of its affinity for water and is commonly

used to adsorb harmful chemicals such as pathogenic viruses, aflatoxin, pesticides, and

herbicides [49, p. 50]. Bentonite is also a common component of drilling muds because

of its impact on the viscosity of the mud.

Montmorillonite also intercalates certain surfactants. It has been shown that both

cationic surfactant molecules [50, p. 367] and anionic surfactant molecules [26, pp. 59 –

61] adsorb on the surfaces of montmorillonite. This fact no doubt has much to do with

the swelling ability of montmorillonite and the ease of accessibility to both the negative

tetrahedral basal plane and the positive exchangeable cations along with the positive

hydroxylated edges of the montmorillonite platelets.

Also, montmorillonite has a cation exchange capability (CEC) of 80 – 150

centimoles of positive charge adsorbed per kilogram of oven-dried clay compared to

kaolinite’s CEC of 2 – 15 cmol/Kg [51]. This high CEC for montmorillonite is

attributable to its high surface area compared to kaolinite. Due to its high internal surface

area (swelling), montmorillonite can have a total surface area of as much as 800 m2/g

while kaolinite, a nonexpanding mineral, may only have a surface area of 10 to 20 m2/g

[47, p. 81 – 82].

Cha, et al., reported on the kinetic and thermodynamic promotion of hydrates due

to clay surfaces, primarily bentonite [41, p. 6494]. Also, Kothapalli showed that in the

presence of biosurfactants, the rate of hydrate formation on bentonite surfaces can change

by as much as four-fold over the rate of formation with no surfactant present.

Page 41: Examination of the Effects of Biosurfactant Concentration

-27-

Rhamnolipid and Emulsan exhibited an increase in rate of hydrate formation of

approximately two-fold while decreasing induction time by 39% and 58%, respectively,

on bentonite surfaces [26, p. 51, p.44].

Nontronite

A related clay surface, nontronite, has a structure similar to montmorillonite as

both are in the smectite classification. The major difference between nontronite and

montmorillonite is that iron is a larger component of the octahedral sheet compared to

montmorillonite. However, aluminum, like iron in montmorillonite, is an impurity within

nontronite. Because of the oxidized iron present in nontronite, the clay appears to have a

greenish yellow or brownish green tint. Nontronite is a slightly harder mineral than

kaolinite or montmorillonite at a Mohr’s hardness of 2 – 2.5. Nontronite has a specific

gravity of 1.7 to 1.8 [48, p. 45]. No apparent studies of gas hydrate formation and

hydrate relationship to nontronite clay existed before this investigation.

Nontronite, while technically a swelling clay, does not contain the same swelling

capacity as its montmorillonite partner as is evident by its relatively minute volume

change when hydrated.

Porous Media, Minerals

Many metal oxides, sulfates, carbonates, phosphates, and other minerals abound

in the earth’s oceans, mantle, and upper crust. Calcite and aragonite are the two most

abundant in rocks and are formed by microorganisms as polymorphs of calcium

carbonate [46, pp. 106 – 107]. Calcite crystallizes in the trigonal system while aragonite

Page 42: Examination of the Effects of Biosurfactant Concentration

-28-

crystallizes in the orthorhombic system [48, p. 46; 52]. The structural pattern of one of

these polymorphs, aragonite, is presented in Figure 2.7.

Figure 2.7. Aragonite Orthorhombic Structure [52]

Aragonite often forms at the Earth’s surface as a precipitate from microorganism

activity and is among the most reactive abundant minerals near the Earth’s surface [46, p.

106]. Aragonite is a white mineral with a Mohr’s hardness of 3.5. As a result of

aragonites precipitation by microorganisms, it commonly appears in shells and other

structural features of sea organisms.

Aragonite, when produced by coral or other organisms, contains an inherent

microporosity. In addition, aragonite is more soluble than its calcite partner, a fact that

often leads to a secondary porosity due to dissolution. Aragonite, being a metastable

compound, also has the ability to alter itself into the more stable calcite form during

diagenesis [53]. To date, no studies related to hydrates exist in relation to aragonite.

Page 43: Examination of the Effects of Biosurfactant Concentration

-29-

-29-

CHAPTER III

THEORETICAL BACKGROUND

Surfactants

Surface-active agents (surfactants) are dual character molecules that have both

hydrophilic and lipophilic tendencies. Because of this dual tendency, surfactants tend to

aggregate along interfacial boundaries [54, p. 5]. If two immiscible liquid layers are in

contact with one another, such as oil and water, then surfactant molecules will spread

along the oil-water interface with its hydrophilic group orienting to the water layer and its

lipophilic group facing the oil layer. If the boundary is a solid surface such as dirt wetted

by water, the surfactant will have much the same effect. Water solvates the hydrophilic

groups while the lipophilic groups surround the dirt particle [54, p. 140 – 143]. It is this

basic principle by which all commercial detergents, soaps, shampoos, and conditioners

function.

Surfactants are generally classified as one of four entities: anionic, cationic,

nonionic, or amphoteric (zwitterionic) [55, p. 3 – 4]. Anionic surfactants create negative

ions (anion) in aqueous solution while cationic surfactants create positive ions (cations)

that are surface active in aqueous solutions. Nonionic surfactants have no functional

groups capable of stabilizing a negative or positive charge in aqueous solution and thus

remain neutrally charged while exhibiting surface-active properties. Anionic surfactants

Page 44: Examination of the Effects of Biosurfactant Concentration

-30-

typically contain a carboxylate, sulfonate, or phosphate group [54, p. 74; 56; 57, p. 8]

while cationic surfactants are typically amine salts [54, p. 74; 57, p. 8; 58]. Nonionic

surfactants commonly have a water-soluble ester group which does not readily donate or

accept protons [54, p. 74; 57, p. 8; 59; 60].

Amphoteric surfactants are a mixture of the previous three classifications,

possessing attributes of each and are commonly called zwitterions in solution.

Amphoteric surfactants have functional groups proficient at stabilizing both negative and

positive charges. When these charges are balanced, the surfactant is neutral like the

nonionic surfactant. When the solution chemistry is shifted, however, these amphoteric

compounds take on either a negative or a positive charge. Amphoteric surfactants

typically carry one of each of the positively and negatively charged functional groups

such as a carboxylate group along with an amine salt [54, p. 74; 57, p. 8]. In this manner,

pH may be manipulated to determine the overall charge of the surfactant molecule [59].

The net charge attained by each surfactant dictates its activity and functionality in

certain situations. An aqueous anionic surfactant in the presence of a positive metal

surface will tend to be attracted to that surface [30, p. 4182]. Conversely, an aqueous

cationic surfactant would be repelled in the same situation. Many times it is also this

nature that dictates how easily a surfactant is stabilized or solvated in an aqueous

solution. Since water is a highly polar solvent, charged surfactants are more easily

dissolved in these situations.

Page 45: Examination of the Effects of Biosurfactant Concentration

-31-

Biosurfactants

Biosurfactants are a specialized class of surfactants that are produced by

microorganisms and are divided into five subcategories listed in the previous chapter [25,

p. 1732]. These subcategories may be anionic, cationic, or nonionic and serve numerous

purposes such as helping to solubilize carbon sources and to aid the bacteria in

transportation through soil media [32, p. 1993; 61, p. 230; 62, p. 60 – 61]. Industrially,

these biosurfactants are becoming more attractive because of their possible exploitation

as emulsifiers, wetting agents, foaming agents, food ingredients, and detergents to name a

few.

Some surfactants and biosurfactants are micelle-forming agents and some are not.

Micelles are highly organized structures consisting of a group of surfactant molecules

aggregating in a specific orientation [55, p. 4 – 15]. Micelles may take on assorted

shapes such as rods, bilayered sheets, vesicles, and worm-like structures [33, p. 569; 63,

p. 1360; 64, p. 3816]. The most typical example of a micelle is the spherical micelle in

which the hydrophilic heads of multiple surfactant molecules align along the surface of a

sphere with their hydrophobic tails directed inward excluding water from the interior of

the sphere. In this manner, micelles are proficient at solubilizing organic substances in an

aqueous solution in which they are typically immiscible.

It is generally understood that cationic surfactants inhibit gas hydrate formation

due to agglomeration retardation and that anionic surfactants promote gas hydrate

formation, possibly through a structuring of water and/or through an increase of gas

solubility [16, pp. 4982 – 4983; 26, pp. 68 – 73; 30, p. 4175; 39, p. 9; 65, p. 53].

Page 46: Examination of the Effects of Biosurfactant Concentration

-32-

Rhamnolipid

A model micelle-forming biosurfactant molecule is the rhamnolipid molecule

produced by the Pseudomonas aeruginosa bacteria. Rhamnolipid is an anionic

biosurfactant containing either one or two carboxylated sugar heads along with one or

two lipid tails. For this reason, rhamnolipid is also classified as a glycolipid.

Rhamnolipid may exist in its micellar form as a sphere, a vesicle, or a lamella (bilayered

sheet) [33, p. 570]. Rhamnolipid’s usual function is most probably to bring carbon

sources into solution where they will be accessible by the bacterial cell. Rhamnolipid

may also aid in transport of carbon sources across the cell membrane by increasing the

cell surface hydrophobicity, a process which can occur at very low biosurfactant

concentrations [66, p. 3262].

Rhamnolipid’s micelle-forming ability is notable as indicated by its relatively low

CMC value. CMC is an important solution property. The CMC is defined as the

concentration at which free surfactant molecules dissolved in solution begin to self-

associate into structure micelles. The CMC is marked by the point of sharp transition in

the surface tension versus concentration curve where the steeply declining shape abruptly

changes to a nearly flat straight line. Micelles are essential because they give surfactants

and biosurfactants their unique activities [67, p. 1229]. Micelles bring the immiscible

organic and water layers together, stabilizing one within the other.

Rhamnolipid was chosen as a test subject because of its demonstrated ability to

effectively promote gas hydrate formation and because of the identification of

Pseudomonas aeruginosa cellular material near areas containing gas hydrates in the Gulf

of Mexico [4, pp. 5146 – 5148]. The method by which rhamnolipid promotes hydrates

Page 47: Examination of the Effects of Biosurfactant Concentration

-33-

has been speculated as a micellar phenomenon [26, pp. 78 – 81]. Not only does a

rhamnolipid solution solubilize natural gas more efficiently, it also lowers the water

surface tension to negate somewhat the large capillary effect of water diffusing through

porous media. By easing the capillary effect, transport of gas molecules and hydrate

clusters are much easier through small pores.

Rhamnolipid has been reported to effectively increase the rate of hydrate

formation in such porous media as sand, bentonite, and kaolinite at high concentrations

[26, pp. 50 – 54]. No data currently exists for hydrate formation with rhamnolipid at

concentrations below 1000 ppm.

Emulsan

Emulsan is a non-micellar biosurfactant and was investigated for its contrast to

the rhamnolipid molecule. Emulsan is a large molecule classified as a polyanionic

biosurfactant in the subclassification polysaccharide-lipid complex. Emulsan has a long

polysaccharide backbone with numerous hydroxyl and carboxylate groups attached.

Because of its many lipid side chains, Emulsan has the ability to spread out over

an oil surface with its lipid chains directed to the organic phase and its polysaccharide

backbone directed to the aqueous phase. In a solely aqueous environment, the Emulsan

molecule bundles up into a coiled structure with the lipid chains directed inward in order

to minimize their contact with the polar water molecules [68]. Along a hydrophobic cell

surface, the Emulsan molecule can attach and extend, acting to direct complex organic

food sources into the cell.

Page 48: Examination of the Effects of Biosurfactant Concentration

-34-

Previous work with Emulsan showed the molecule to be proficient at directing

hydrate formation on surfaces such as kaolinite and, to a much greater extent, bentonite

[26, p. 73]. No data exists of Emulsan-directed hydrate formation at concentrations less

than 1000 ppm.

Adsorption Theory

Adsorption is defined as adhesion to a solid surface or body by a gas, solute, or

liquid in an extremely thin layer of molecules. Many times in the chemical industry,

adsorption is used to describe a catalytic process between a gas phase and a solid catalyst

such as the use of platinum and rhodium catalyst to reduce nitrogen oxide compounds

(NOx) from automobile exhausts. In soil sciences, the term adsorption commonly refers

to a process by which an organic material or an ion adheres to a soil surface. Sloan also

uses an analogy of adsorption to describe the process by which gas is enclathrated into

gas hydrates [3, pp. 208 – 211].

Adsorption Classification

Adsorption is a net result of two distinct interactions. Interactions occur between

the molecule being adsorbed (adsorbate) and the surface on which the molecule is

adsorbed (adsorbent.) Interactions can also occur between two adsorbate molecules. The

relative strength of the adsorbate-adsorbent and the adsorbate-adsorbate interactions

ultimately determine the type of adsorption that transpires. Adsorption interactions are

divided into four categories: C-type, S-type, L-type, and H-type. Examples of the four

types of adsorption are presented graphically in Figure 3.1 [69, pp. B-277 – B-278].

Page 49: Examination of the Effects of Biosurfactant Concentration

-35-

C - Type S - Type

H - TypeL - Type

Qua

ntity

Ads

orbe

d

Concentration

Figure 3.1. Adsorption Isotherms [69, p. B-278]

C-type adsorption, often called “constant partitioning”, is described most

effectively by a plot of adsorbate concentration versus amount of adsorbate adsorbed

called an adsorption isotherm. When these quantities are plotted for a C-type isotherm,

the result is a straight line. The shape of this isotherm tells a very important fact about C-

type adsorption. The relative affinity of adsorbate molecules for the adsorbent is constant

[69, p. B-277]. More importantly, this type of adsorption is not dependent on any type of

bonding between the adsorbate and adsorbent. C-type adsorption, like most adsorption

isotherms, is only reliable at low concentrations. This type of adsorption is usually

associated with nonpolar organic molecules.

Page 50: Examination of the Effects of Biosurfactant Concentration

-36-

S-type adsorption, sometimes called cooperative adsorption, exists when

adsorbate-adsorbate interactions are dominant over adsorbate-adsorbent interactions. The

S-type isotherm is an S-shaped curve which shows little or no change at very low

concentrations, rises sharply at intermittently low concentrations, and then levels off to a

constant value at higher concentrations. The slow increase at very low concentrations of

adsorbate indicates that clustering or agglomeration of adsorbate molecules is preferred

over adsorption [69, p. B-277].

The final two types of adsorption, L-type and H-type, are both types of

chemisorption and are associated with chemical bonding rather than physical attractions.

These types of isotherms are referred to as Langmuir isotherms, the H-type being an

extreme case of the L-type. These isotherms are characterized by a high degree of

adsorption at very low concentrations of adsorbate. The Langmuir isotherm and H-type

isotherm are described by the Langmuir and Freundlich equations presented below [69, p.

B-278].

+

=i

ii Kc

Kcbq

1 (3.1)

Here, the amount of adsorbate adsorbed qi is related to the equilibrium concentration of

the adsorbate in solution ci through the affinity of the adsorbate K (which determines the

slope of the isotherm) and the maximum adsorption capacity b. As stated before, when K

is very large, the L-type isotherm is typically referred to as an H-type isotherm.

L-type and H-type isotherms also inherently make five key assumptions. (1) The

surface of interest has a finite number of identical sites for adsorption, each having a

capacity for one molecule only. (2) Adsorption is a reversible process. (3) Adsorbate

Page 51: Examination of the Effects of Biosurfactant Concentration

-37-

molecules cannot move laterally between absorption sites. (4) The energy of adsorption

is the same for every molecule and every site independent of surface coverage. (5)

Interaction between adsorbate molecules is negligible [70].

Adsorption in Soils

Adsorption in soils is usually associated with cation or metal ion exchange.

Remediation of heavy metal contaminated soils requires an innate understanding of soil

adsorption principles where lead and mercury may be leached into the ground.

Adsorption of heavy metal ions is usually of the L-type or H-type because of their high

affinity for the negatively charged surfaces which abound in soils. Also, dissolved

organic matter (DOM) may be problematic in soil remediation typically being of the S-

type adsorption. DOM has a low affinity for soil surfaces because it typically has a low

polarity or no polarity.

Valence of cations plays a paramount role in the replaceability of a cation for

another cation. This trend referred to as the lyotropic series indicates that diavalent

cations such as calcium and magnesium are adsorbed tighter to soil surfaces than are

monovalent cations such as sodium or potassium [47, pp. 150 – 151]. In addition, cations

with larger dehydrated radii tend to be retained more proficiently than cations of the same

valence with smaller dehydrated radii.

Anionic adsorption in soils usually takes place at surfaces with a high degree of

hydroxyl sites and many times involves a loss of water by the anion being adsorbed [69,

pp. B-285 – B-287]. Adsorption of anions may be either an inner-sphere process or

outer-sphere process. Inner-sphere processes are strong surface adsorptions where water

Page 52: Examination of the Effects of Biosurfactant Concentration

-38-

plays no role in the adsorption (i.e. the adsorbate is chemically bound to the adsorbent.)

Outer-sphere processes involve the mediation of water by solvation of the anion and are

necessarily weaker attractions [69, pp. B-241 – B-242].

Outer-sphere processes should dominate the adsorption of anionic biosurfactant

molecules on sand and kaolinite since sand and kaolinite have a high degree of oxide

surfaces with relatively few hydroxyl surfaces. In contrast, inner-sphere processes should

dominate the adsorption of anionic biosurfactants on the surfaces of bentonite and

nontronite because of their distinct hydroxyl edge effect. Bentonite gains an added

advantage in this sense, as it will also intercalate anionic molecules into its cation-

stabilized interlayers more so than nontronite. Similarly, the surface of aragonite should

be conducive to inner-sphere adsorption with its abundant Ca+2 sites.

In addition, the possibility remains for anion repulsion or negative adsorption.

Anions with high charge densities such as Cl-, NO3-, and SO4

- are typically repulsed in

negatively charged soil surfaces. This negative adsorption can often lead to a higher

concentration of the anion in solution after being introduced to the soil. This observation

comes about by two processes. 1) The anions are expelled from the diffuse double layer

and concentrated into the bulk fluid. 2) Hydration of the soil surface removes water from

the bulk solution thus increasing the anion concentration [47, pp. 172 – 173].

Adsorption and Gas Hydrates

Gas hydrate formation is an interfacial phenomenon. Gas hydrates form most

proficiently at the source of highest gas to water ratio. Therefore, solubility of gas into

the liquid water phase is essential in increasing hydrate formation. The other key

Page 53: Examination of the Effects of Biosurfactant Concentration

-39-

component of hydrate formation is the statistical probability of a hydrate cage forming in

proximity to a guest molecule. Biosurfactants seem adept at increasing gas solubility.

Beyond this, they also seem proficient in some cases at directing, through adsorption, the

construction of the hydrate lattice thereby bringing the guest molecule into the right place

at the right moment. In this sense, biosurfactants seem to be most effective catalysts for

hydrate formation.

Induction Time & Nucleation Theory

Induction time is a temporal measure of the supersaturation of a gas hydrate

mixture. Induction time is the time required for gas hydrates to form a critical nucleus

cluster after achieving saturation and begin rapid hydrate precipitation. The form and

formation rate of clathrate compounds, like other crystalline compounds, are dictated by

nucleation theory. Vysniauskas and Bishnoi presented in 1983 a three-step mechanism

for the nucleation of gas hydrates: initial clustering, critical size nucleation, and

propagation and crystal growth [71, p. 1069].

The onset of gas hydrate formation can be viewed as a random ordering of

individual water molecules into a three-dimensional, hydrogen bonded system known as

the “network-cluster” model [3, pp. 68 – 74]. When approaching the freezing point,

water molecules form and break hydrogen bonded polygon structures at random. If the

concentration of gas is high enough in this dynamic system, then these hydrogen-bonded

structures will form around a gas molecule, and the gas will be encapsulated in a cell

structure. Quite intuitively, this statistical probability is greatest at the gas-water

Page 54: Examination of the Effects of Biosurfactant Concentration

-40-

interface, which has lead many to conclude that hydrate formation is an interfacial

process.

Vysniauskas and Bishnoi (1983) contended that the initial stage of hydrate

formation was followed by a period where these well-dispersed hydrate monomers

encountered other water molecules and other guest molecules. The hydrate structure

would then grow by the two-body interaction process until a critical size was achieved.

This critical size, known as a critical cluster, would then pass a point of irreversible

rearrangement and inherent stability.

The final step in the nucleation process is growth of the critical nucleus closely

followed by precipitation of gas hydrate “polymers.” This step is a rapid one and easily

visualized both physically and experimentally. It is also accompanied by a large

liberation of latent heat, as hydrate formation is an exothermic process. The overall

mechanism is described in the following equations.

( ) ( )yy OHMOHOHM 2122 ...↔++ − (3.2a)

( ) ( )cz OHMOHMOHM 222 .↔++ (3.2b)

( ) ( )nm OHMOHMOHM 222 .↔++ (3.2c)

In this formulation, M is the guest molecule and y, z, m, and n denote the size of any

given cluster. The c subscript denotes that at this point, the clathrate has assumed a

critical size. The ellipsis in Eq. 3.2a denotes a physical attraction while the single period

in Eq. 3.2b and 3.2c denote a bond via van der Waals forces.

In cases of some simple hydrates, an induction period is absent indicating that the

formation of a critical nucleus occurs rapidly [3, pp. 85 – 90]. However, in most other

Page 55: Examination of the Effects of Biosurfactant Concentration

-41-

cases, the induction period can be significant. The presence of an induction period

suggests that a rate-limiting process is also present. Since a critical nucleus is

thermodynamically stable (or metastable) and the propagation step is rapid, Equation 3.2a

must contain the rate-limiting step. This step is thermodynamically unstable and

statistically random.

Curiously, studies have shown an apparent memory effect of water upon hydrate

formation and dissociation and even upon freezing and melting [71, pp. 1064 – 1065].

Upon thawing ice or dissociating gas hydrates, the hydrogen-bonded water lattice is not

simply destroyed. Rather, the basic structure seems to be preserved while essential cage

breakage does occur to allow the gas to escape in the case of hydrates. This statement is

supported by data showing that, upon reformation with dissociated hydrate water or with

thawed water, induction time is extremely short and sometimes missing altogether.

Gas Hydrate Formation Rate (Kinetics)

Vysniauskas and Bishnoi also proposed a semi-empirical rate formulation based

on their proposed mechanism [71, pp. 1069 – 1071]. The related their mechanism to

reaction kinetics assuming that the rate of hydrate formation must be proportional to the

concentrations of water, gas, critical nuclei, and the interfacial surface area as through an

Arrhenius rate constant kr.

[ ] [ ] [ ]qnc

msr MOHOHakr 22= (3.3)

After making substitutions for the Arrhenius rate constant and substituting the

Boltzman distribution function for the critical size cluster along with some empirical

assumptions, they arrived at a rate equation for hydrate growth that is mostly theoretical.

Page 56: Examination of the Effects of Biosurfactant Concentration

-42-

The rate equation was a function of interfacial area as, activation energy ∆Ea, temperature

T, pressure P, the degree of subcooling ∆T, a lumped pre-exponential factor that

encompassed heat and mass transfer effects A, and two empirical parameters a and b.

Their rate equation is presented as Equation 3.4.

γPTa

RTE

Aar ba

s ⋅

∆−

∆−= expexp (3.4)

Equation 3.4 was derived for a stirred system where the gas-water interface is

renewed due to agitation. In quiescent systems the process is somewhat different but the

same parameters determine hydrate formation rate. If the interfacial surface is stagnant, a

thin film of gas hydrates forms on the surface of the liquid and blocks gas transfer to the

liquid below. This film effectively terminates further hydrate formation by reducing

mass transfer rates (A essentially approaches zero.) The only manner of hydrate

propagation in this instance is through gas diffusion through hydrate capillaries which is

an extremely slow process, sometimes taking weeks.

Past work has shown that the introduction of a surface-active agent can renew the

interfacial area of a quiescent system by 1) reducing the capillary forces, 2) increasing

mass transfer through gas dissolution, and 3) ordered packing of hydrate clusters away

from the gas-water interface [30, pp. 4177 – 4185]. Recently, others have shown that

Vysniauskas and Bishnoi’s rate equation holds approximately true for quiescent systems

under the influence of biosurfactants [26, pp. 56 – 59].

Page 57: Examination of the Effects of Biosurfactant Concentration

-43-

Phase Equilibria

Phase equilibria diagrams for hydrate systems give valuable information about the

conditions under which hydrates should form. Before the principles of nucleation and

induction (kinetic properties) can be fully understood, the principles of where the

thermodynamic equilibrium curves lie need to be established for reference. Two early

methods for theoretical determination of the three-phase equilibrium line for a water-gas-

hydrate system were developed, the gas gravity method and the distribution coefficient

method, and are still in frequent use today. However, the most accurate methods

employed today are statistical approaches.

Gas Gravity

The gas gravity method has been employed since 1945 to roughly estimate the

onset of hydrate formation or dissociation [3, pp. 136 – 139]. The gas gravity value is

analogous to a liquid’s specific gravity. Gas gravity (G.G.) is calculated by dividing the

molecular weight of the hydrate forming gas by the molecular weight of air as shown in

Equation 3.5. The subscript i denotes a component of the gas mixture.

Air

n

iii

MW

MWyGG

∑=.. (3.5)

The principle behind this method is that gases, irrespective of their components,

experience a shift in equilibrium on a P-T curve based on their gas gravity. Gases that

are lighter than air (G.G. < 1) are shifted up the P-T chart from the base gravity value of 1

with little or no change in the slope of the curve.

Page 58: Examination of the Effects of Biosurfactant Concentration

-44-

The gas gravity method has some basic generalization flaws. Namely, the

correlation was developed for hydrocarbon gases, and unacceptable errors could occur

when extrapolating data containing non-combustibles. The gas gravity method is a good

approximation for a simple initial guess but more accurate methods for determining phase

equilibria now exist.

Distribution Coefficient Method

The distribution coefficient method was conceived by Wilcox, Carson, and Katz

in 1941 and is the precursor to statistical thermodynamic methods [3, pp. 144 – 159].

Wilcox et al. hypothesized that hydrates could effectively be viewed as vapor-solid

mixtures instead of a three-phase distribution containing water. By effectively

eliminating the presence of water in their calculations, they arrived at a simple vapor-

solid distribution coefficient formulation shown in Equation 3.6.

si

ivsi x

yK = (3.6)

In Equation 3.6, yi is the mole fraction of gas component i in the water-free vapor phase

and xsi is the mole fraction of gas component i in the water-free solid.

The practicality of this formulation is that the relative ability of a gas component

to form hydrates is easily seen. Those gas components that yield a vapor-solid

distribution coefficient greater than one preferentially remain in the gas phase, and those

components with a Kvsi value less than one easily form hydrates.

Also, equilibrium temperatures and pressures can be calculated by this method

through an 18-parameter empirical formula relating the natural log of Kvsi to temperature

Page 59: Examination of the Effects of Biosurfactant Concentration

-45-

and pressure. With today’s increased computer calculating power, the Wilcox,Carson,

and Katz method serves as a quick and accurate estimation tool not limited to only

hydrocarbon gases.

Statistical Thermodynamics

Statistical thermodynamic determination of gas hydrate equilibrium relies heavily

on the Langmuir adsorption analogy. In this analogy, five major assumptions are made

for single component systems [3, pp. 209 – 210]:

1) The enclathration of gas molecules occurs at discrete cavities on the surface.

2) The energy of enclathration on the surface is independent of the presence of

other enclathrated molecules.

3) The maximum amount of enclathration corresponds to a monolayer, or one

molecule per site.

4) The enclathration is localized and occurs by collision of gas phase molecules

with vacant cavities.

5) The declathration rate depends only on the amount of material on the surface.

This analogy allows us to determine the fractional filling of gas hydrate cavities by

comparing it to the established Langmuir equation with slight modifications.

∑+=

JJJi

kkiki PC

PCY

1 (3.7)

In Equation 3.7, the fraction Y of i cavities filled by molecule type k can be related

to the pressure of the system P and a set of Langmuir constants. If this equation were

used for simple hydrates, Cki would be the only unknown parameter in this equation.

Page 60: Examination of the Effects of Biosurfactant Concentration

-46-

Overall Mechanism

Determination of an overall mechanism of gas hydrate formation in porous media

with the presence of biosurfactants is a lofty goal. However, clues as to the true nature of

the mechanism can be seen in parts of the governing equation of each interaction (simple

hydrate phase equilibrium, biosurfactant-gas interaction, biosurfactant-porous media

interaction, hydrate-porous media interaction, etc.) Based on these interactions, a

seafloor mechanism for hydrate formation can be inferred.

To say that hydrate formation is simply an interfacial phenomenon is to imply that

all other interactions between surfaces and dissolved species are negligible. This,

however, is not the case. Hydrate clustering followed by nucleation followed by crystal

growth must be a viable mechanistic pathway, but how are these steps altered by the

presence of other species? It is contended here that both biosurfactants and porous media

interact with hydrate nucleation in a synergistic fashion to promote hydrate formation,

inhibit hydrate formation, or leave hydrate formation unchanged.

Biosurfactants must by their dual nature both solubilize hydrocarbon gas and

reduce the capillary effect of water by lowering the surface tension. These effects should

act to increase the overall mass transfer rate of gas past the interfacial boundary along

with structuring the water interface very near the micelle surface and allow gas hydrates

to form in the bulk fluid more readily. In addition, the surface tension lowering effect of

the biosurfactant should improve the ability of water to be carried to the water-gas

interface through capillary effects. Capillaries of adequate size within porous media or

Page 61: Examination of the Effects of Biosurfactant Concentration

-47-

with hydrate structures themselves should imbibe water easily due to the lower surface

tension of the solution relative to distilled water [72, p. 589].

As a result, the typical hydrate sheet barrier of quiescent systems must be absent

allowing more hydrates to form at a faster rate. This effect may be attributable to micelle

formation but more likely is some function of particle size (whether it be a micelle or a

large molecule such as Emulsan) and solubility.

It would also be known that surfaces such as bentonite clay promote hydrate

formation while others such as kaolinite do not to an appreciable extent. Evidence has

been provided for the hydroxyl edges of bentonite or montmorillonite to be the active

agent for this increase in formation rate. If this is truly consistent, then other surfaces

with similar properties such as nontronite should have similar effects on hydrate

formation. It would also be reasonable to assume that surfaces such as sand would not

promote hydrates to any extent.

There must also be an important interaction between the biosurfactant and the

porous media. Soil adsorption of organic materials is common and expected especially in

the case of an anionic surfactant. Adsorption between the anionic surfactant is expected

to occur at the hydroxyl edges rather than at the tetrahedral oxide planes common in

many types of sediment. Therefore if adsorption were to occur, it would be most

dominant in the smectite clays and most prominent in high swelling clays such as

montmorillonite where adsorption into the interlayers seems most plausible.

Water structuring is also expected to be salient along the basal planes of

sediments, more so in sediments containing hydroxylated sites due to hydrogen bonding.

Page 62: Examination of the Effects of Biosurfactant Concentration

-48-

The structuring of water into hydrate-shaped cages as suggested by Cha, et al. would

lower the Gibb’s free energy needed to form the initial clusters of hydrates [41, p. 6494].

A general feasible mechanism in the case of micellar anionic surfactants is

hypothesized as the following. As concentrations increase and micelles become

dominant, the increased solubility, the decrease in the capillary effect, and the possible

increase in nucleation sites due to small micelles all serve to increase the effective

interfacial area and thereby increase hydrate formation markedly over the case of no

surfactant. In the presence of sediment, the micellar effect can be either accentuated or

depressed depending on the types of interactions between the biosurfactant and the

surfaces. Bentonite and nontronite should amplify the effect of the biosurfactants

because of their high structuring effects associated with the hydroxyl edges while Ottawa

sand and kaolinite should not change appreciably.

For the case of the non-micellar, polyanionic Emulsan, the effect should be

somewhat different. The Emulsan should solubilize natural gas to a lesser extent due to

the absence of discreet micelles. The large molecular weight and low solubility of

Emulsan should however make it a good prospect for a nucleation site. As a result, the

effect of Emulsan to increase formation rate should be more of a function of particle size

and quantity than solubility and increase of effective interfacial area. In this manner, the

Emulsan should receive much coupled effect of sediment interaction because less hydrate

formation should take place in the bulk fluid where the sediment surfaces would be

accessible. Also, the tertiary structure of Emulsan should affect the rate of hydrate

formation since water structuring near the Emulsan molecule seems plausible. When

coiled or uncoiled, the molecule would have different interactive properties.

Page 63: Examination of the Effects of Biosurfactant Concentration

-49-

The above mechanism relies on a number of interactions that may or may not be

differentiable or discernible. The focus of this paper, particularly the following chapters,

is a qualitative explanation of experimental results. Some points are made very clearly

by the data and others must be suggested or extrapolated. But in each case, every effort

is made to relate the data to theory-grounded molecular interactions.

Page 64: Examination of the Effects of Biosurfactant Concentration

-50-

-50-

CHAPTER IV

EXPERIMENTAL METHODS

Hydrate Experimental Setup

The experimental method consisted of examining the interaction of two dissimilar

biosurfactants (one which forms micelles and one which does not) with primarily five

porous media surfaces. Three of these surfaces were examined in detail while the others

were compared at points of extremity. The porous media of concern were purified

Ottawa sand, bentonite clay, kaolinite clay, nontronite clay, and aragonite (CaCO3.) The

biosurfactants evaluated were rhamnolipid, a micelle-forming surfactant from the bacteria

Pseudomonas aeruginosa, and Emulsan, a high molecular weight bioemulsifier from the

species Acinetobacter calcoaceticus.

A sample cup was constructed to simulate sediment/natural gas/water interaction

on the ocean floor and to minimize the hydrate formation time. The sample cup was

designed for maximum heat and mass transport to the sand/clay pack that it contained,

such that heat and mass transfer effects would not be rate limiting. Each sample

container consisted of a 50 ml polypropylene cup with a 2-inch long section of ½-inch

diameter clear PVC pipe epoxied in the center of the bottom of the cup. This

configuration was chosen to give a thin annular space for sediments where gas could

access the sediment from two directions. Ports were drilled into the side of the

Page 65: Examination of the Effects of Biosurfactant Concentration

-51-

polypropylene cup and the clear PVC to increase mass transport of gas to the porous

media. One-sixteenth inch diameter holes were drilled around the perimeter of the cup

and piping every 45 degrees. Vertically, the ports were placed ½-inch apart and

staggered ¼-inch. Five ports were also drilled into the bottom of the sample cup in the

inner PVC ring. These ports allowed for excess water drainage. A schematic of the final

sample cup is shown in Figure 4.1.

2"

2"

7/8"

1/4"

1/2"1/2"

45o

Figure 4.1. Diagram of Hydrate Sample Cup (Drawn to Scale) The reaction vessel was a 450-mL pressure cell manufactured by Parr Instrument

Company. The vessel was a Model 4762 constructed of 316 stainless steel and rated for a

Page 66: Examination of the Effects of Biosurfactant Concentration

-52-

maximum pressure of 2950 psi. The cell was 2.5 inches inside diameter by 5.94 inches

tall. It was fitted with two 3-wire platinum resistance temperature devices (RTDs) and an

Omega, Inc. pressure transducer calibrated from 0 – 500 psig. One RTD probe was

placed at the top of the vessel through one of the 7/8th-inch female nominal pipe thread

(FNPT) ports to measure the temperature of the bulk gas. The other RTD was placed just

at the surface of the sample through a 9/16th-inch branched fitting (illustrated in Figure

4.2) to measure the heat liberated upon hydrate formation.

Figure 4.2. Hydrate Formation Cell Photograph

RTD Probes

Relief Valve

Pressure Transducer

Parr Reaction Vessel

Natural Gas Inlet

Page 67: Examination of the Effects of Biosurfactant Concentration

-53-

The vessel was fitted with a pressure relief valve set to relieve at 500 psig which was also

located on the 9/16th-inch branch fitting. A diagram of the reaction vessel is shown in

Figure 4.3.

Cylinder, 450 mL,5.94" deep

Cylinder Head

Drop band with setscrew

Compression Ring

Split ring, pair, withcap screws

Gasket, PTFE

9/16" Cap screws,304 SS

Cross-sectional View

Top View

Drop band withset screw9/16" FNPT

7/8" FNPT

Figure 4.3. Diagram of Hydrate Reactor Vessel (Not to Scale)

Page 68: Examination of the Effects of Biosurfactant Concentration

-54-

The entire pressure vessel containing the packed media sample was submerged in a

constant temperature water bath for cooling.

Hydrate Preparation

Before preparing the porous media sample, all Ottawa sand to be used was

cleaned with ethanol to remove particulate matter. Approximately 50 ml of sand was

placed in a 150 ml beaker with enough ethanol to reach 120 ml total volume. The sand

was then sonicated with an ultrasonic horn for two minutes at 200 watts and 20 kHz

before the ethanol was drained along with any residue. This procedure was repeated and

the cleaned sand set under the hood overnight to thoroughly dry. The top of the beaker

was covered but not sealed with aluminum foil to prevent new particulate matter from

settling on the sand while still allowing the sand to dry. The sand was then placed in a

vented oven overnight at ~60oC to ensure that the sand was free of all volatile organic

material.

First, to prepare an individual sample, approximately 21 g of ethanol-cleaned

Ottawa sand was placed in the annulus of the sample cup. Second, a 1.25 g layer of the

sediment of interest was coated on top of the sand layer. These two steps were then

repeated so that four distinct layers resulted. Finally, a 25 g layer of Ottawa sand was

placed on top of the fourth layer, leaving a thin semicircular recess around the upper rim

of the sample cup. A thin 0.5 g layer of subject porous media was then laid in this recess

before proceeding to the soaking process. A schematic of the filled sample cup is shown

in Figure 4.4.

Page 69: Examination of the Effects of Biosurfactant Concentration

-55-

Ottawa Sand

21 g

21 g

25 g Clay or Carbonate

Cross-sectional View

1.25 g

0.5 g

Ottawa Sand

Clay or Carbonate

Top View

Figure 4.4. Filled Sample Cup The perforated sample cup was placed in a 150 ml polypropylene container. The

desired surfactant solution was prepared by weighing surfactant on a Mettler balance of

accuracy ±0.0001 g and mixing the surfactant thoroughly with an appropriate amount of

distilled water (or seawater.) The surfactant solution was then poured around the sample

cup letting the solution diffuse into the pack through the 1/16th-inch ports. The container

Page 70: Examination of the Effects of Biosurfactant Concentration

-56-

was capped and allowed to soak for approximately 30 minutes, after which the sediment

pack was removed from the water, placed on an absorbent cloth, covered, and allowed to

drip drain for another 30 minutes.

The drained sample was placed in the Parr reaction vessel and sealed. Air was

immediately purged from the vessel with a gas mixture of 90% methane, 6% ethane, and

4% propane before pressurizing to slightly above 320 psig. The vessel was brought to

20oC by submerging it in a 5-gallon reservoir fitted with a copper heating/cooling coil

and maintained at constant temperature with a Cole Parmer circulating bath. As needed,

the vessel pressure was adjusted to consistently reach the desired 320 psig at 20oC. The

Parr reactor containing the sample was then abruptly submerged into the constant

temperature water bath at 0.5oC.

The temperature of the hydrates, the temperature of the bulk gas, and the pressure

of the vessel were recorded at 30-second intervals. The surface tension of the

biosurfactant solution was also measured with the help of a model ST-PLUS surface

tensiometer from Tantec, Inc.

Adsorption Preparation

For each condition of hydrate formation, a series of tests were run to determine

the extent of adsorption of the particular biosurfactant on the media being tested.

Quantifying the amount of biosurfactant removed from the bulk solution by the specific

sediment substantiated this extent of adsorption. This removal is attributable to

adsorption on the surface of the porous media. However, in the case of Emulsan, some of

Page 71: Examination of the Effects of Biosurfactant Concentration

-57-

the biopolymer lost from solution may be due to precipitation and settling, as Emulsan is

a large molecule capable of agglomerating and settling out of solution.

To test for adsorption, a vertical packed-column configuration was employed

where biosurfactant solution contacted the porous media. The adsorption column was

constructed from a 2 feet long section of 1-inch diameter polyvinyl chloride (PVC) pipe

capped at the lower end and open at the top. The capped end of the column was tapped

and threaded to accommodate a ¼-inch polypropylene nozzle. Tygon tubing was

attached to the nozzle and clamped for control of the effluent from the column. Also, a

stainless steel wire mesh screen (greater than 30 mesh) was inserted into the base of the

column to prevent sand from fluidizing and being carried through the column. Adsorption

(or precipitation) was inferred by measuring the surface tension of the surfactant solution

before and after it passed through the column.

Each hydrate sample was contacted by approximately 250 milliliters of surfactant

solution for approximately 30 minutes before removal from the “soaking” solution and

drainage prior to hydrate formation. To maximize consistency of the results, 250

milliliters of surfactant solution at 20oC were also used for the adsorption tests.

For each adsorption test, the column was calibrated to compensate for any effects

of PVC and wire meshing on adsorption. First, surfactant solution was mixed to the

appropriate concentration, measured for surface tension, and then poured into the open

end of the vertical adsorption column. The surfactant was allowed to permeate the

porous media for approximately 30 minutes with the Tygon tubing on the effluent line

clamped. This period allowed ample time for the biosurfactant to establish adsorption

equilibrium with the PVC walls and wire meshing.

Page 72: Examination of the Effects of Biosurfactant Concentration

-58-

After the customary 30 minutes, 100 milliliters of the solution were drained from

the bottom of the column and discarded to eliminate any concentration gradients that

might have been created by the wire mesh. Then, three separate 20 ml samples of

surfactant were extracted from the bottom of the column and measured for surface

tension. If any variation between initial and final surface tensions of the solution

occurred at this point, it could be directly attributed to the PVC and/or wire meshing

adsorption of biosurfactant.

After completing the calibration, the column was rinsed thoroughly and dried. To

the top of the column, 70 ±0.05 g of ethanol-cleaned Ottawa sand was added along with

another 250 ml of surfactant solution. The calibration procedure was repeated, and three

20 ml samples of the effluent surfactant solution were collected and tested for disparity in

surface tension. This procedure was reiterated for concentrations of 10 parts per million

(ppm), 100 ppm, and 1000 ppm of both rhamnolipid and Emulsan solutions. Five porous

media combinations were tested. The porous media tested were 70-g of Ottawa sand and

a 67-g/3-g mixture of Ottawa sand/mineral with mineral being bentonite, kaolinite,

nontronite, or aragonite. If any residual mineral existed in the sample after removal from

the column, the sample was centrifuged for 20 minutes at 6000 rpm to remove the

sediment before surface tension measurements were taken.

Equipment

Mass Balance

The balance used for all mass weighing was an AG285 model Mettler Toledo

balance purchased from Mettler-Toledo, Inc. of Columbus, OH. For weights ranging

Page 73: Examination of the Effects of Biosurfactant Concentration

-59-

from 0 to 41 g, the instrument has an accuracy of ±0.01 mg with a repeatability of ±0.02

mg. For heavier weights, the balance has a maximum allowable weight of 81 g and an

accuracy of ±0.01 mg with a repeatability of ±0.05 mg.

Constant Temperature Bath

The refrigerated, constant-temperature bath used for hydrate formation was a

Model RTE-17 circulating bath purchased from Thermo NESLAB of Newington, NH.

The bath has a temperature range of –22oC to +150oC at a temperature stability of

±0.01oC. The unit has a non-CFC air-cooled refrigerating system, a circulating pump,

and a 4.5-gallon stainless steel bath basin (this basin allowed the test cell to be fully

submerged in the coolant fluid.) Distilled water was used as the circulating medium.

Equilibration Bath

The circulating bath used to establish the sample at 20oC before introduction into

the cooling bath was a Model 9105 purchased from Cole Parmer. The temperature bath

had a range of –20oC to 150oC and contained a six-liter stainless steel reservoir. The unit

had a temperature stability of ±0.05oC with a readout accuracy of ±0.5oC. The inlet and

outlet of the bath’s 15 l/min circulating pump were connected to a ¼-inch diameter

copper tubing coil submerged in a five-gallon water reservoir.

Sonicating Horn

The ultrasonic generator used for Ottawa sand cleaning was a 500-watt Model

VC501 unit manufactured by Sonics and Materials, Inc. The unit converted 60 Hz line

voltage to 20 kHz of electrical energy. The amplitude of the frequency was kept at 40%

Page 74: Examination of the Effects of Biosurfactant Concentration

-60-

of the maximum amplitude to prevent damage to the sonic horn. The converter was a

CV26 model attached to a standard ½-in horn. The duration of the ultrasonication was

two minutes.

Surface Tensiometer

The surface tensiometer used to measure the surface activity of the surfactant

solutions was a ST-PLUS model unit purchased from Tantec, Inc., of Schaumburg, IL.

The unit is capable of Wilhelmy Plate, Wilhelmy Plate with Detach, and DuNouy Ring

measurement methods. The Wilhelmy Plate with Detach method was employed for all

measurements. The tensiometer had a measurement range of 0 to 100 mN/m and an

accuracy of ±0.01 mN/m.

pH meter

A Model 620 pH meter was purchased from Thermo Orion of Beverly, MA to

determine pH range of surfactant solution. The instrument was connected to a Model

6165 Sure-Flow solid-state probe also purchased from Thermo Orion. The unit had a pH

range of 0.00 to 14.00 with a relative accuracy of ±0.01.

Reaction Vessel

The reaction vessel for hydrate formation was a Model 4762 stainless steel

pressure vessel purchased from Parr Instrument Company of Moline, IL. The pressure

container held 450 ml of volume and was rated for a maximum working pressure of 2950

psi at 350oC. The maximum working temperature was limited to 350oC because of the

Page 75: Examination of the Effects of Biosurfactant Concentration

-61-

PTFE gasket sealing the vessel. The head of the vessel had two 7/8-inch FNPT ports and

one 9/16-inch FNPT port.

Sample Container

The sample container for hydrate formation was a 50-mL polypropylene weighing

cup epoxied to a 2-inch long, ½-inch diameter section of PVC pipe. The construction and

diagram of the sample container are detailed in Figure 4.1.

RTD probes

The RTDs used in these experiments were 3-wire Platinum 100Ω models

purchased from Omega. Both probes were protected by 1/8-inch 304 stainless steel

sheaths. The Diagnostic Instrumentation and Analysis Laboratory (DIAL) at Mississippi

State University calibrated each RTD. The RTDs were calibrated for a range of 32oF to

200oF with comparisons to National Institute of Standards and Technology (NIST)

traceable equipment. The probes had a standard temperature deviation of ±0.3oC at 0oC

or ±0.8oC at 100oC.

Pressure Transducer

The pressure transducer was purchased from Omegadyne, Inc. of Sunbury, OH.

The pressure transducer was a Model PX02C1-500G5T requiring 24 – 32 Vdc excitation

with a pressure range of 0.00 to 500.00 psig. The pressure transducer was also calibrated

by DIAL with comparisons to NIST traceable equipment. The transducer repeatability

and hysteresis were both ±0.05% full-scale output (FSO) and the linearity was ±0.15%

FSO.

Page 76: Examination of the Effects of Biosurfactant Concentration

-62-

Linear Power Supply

A Model U24Y101 linear power supply purchased from Omega supplied the

required 28 Vdc excitation to the pressure transducer. The power supply converted a

110-V AC signal to a 24-VDC signal.

Pressure Relief Valve

The pressure relief valve was an adjustable Swagelok R3A series spring-action

valve with a red color designation. The valve was tested by a positive displacement

pump to relieve at the proper 500 psig.

Data Acquisition System

National Instruments’ FieldPoint (hardware) and LabVIEW (software) products

were used to acquire all data from the experiments. The pressure transducer was

connected to a Model FP-AI-110 analog input board while the two RTDs were connected

to a Model FP-RTD-122 3-wire RTD input board. The computer was connected to a

Model FP-1000 network interface board by a 9-pin serial cable. All three hardware

modules were jumpered to a Model FP-PS-4 24 VDC power supply supported by a 110V

outlet power cord. The data acquisition software LabVIEW was configured with the help

of National Instrument technicians. The FP-RTD-122 has a typical accuracy of ±0.15oC

and a maximum error of ±0.3oC with a resolution of ±0.016oC. The FP-AI-110 has a

resolution of ±190µV and a gain error of ±0.1% FSO.

Page 77: Examination of the Effects of Biosurfactant Concentration

-63-

Digital Camera

All photographs of formed hydrates were taken with a Sony CD Mavica digital

camera with 2.1 Megapixel resolution. The Sony camera wrote digital pictures to a 156

MB, 8 cm rewritable CD drive.

Materials

Rhamnolipid

The biosurfactant rhamnolipid was purchased from the Jeneil Biosurfactant

Company located in Saukville, Wisconsin. The rhamnolipid is a processed 25% active

aqueous solution, which is sterilized and practically free of all proteins. Jeneil markets

this formulated rhamnolipid solution under the product name JBR 425. Both R1

(C26H48O9) and R2 (C32H58O13) types of rhamnolipid are present in JBR 425 to an

unspecified extent.

Emulsan

Emulsan is a bioemulsifier produced by the bacteria Acinetobacter calcoaceticus.

In solid form, it is a fine yellow powder. Emulsan is comprised of a long polysaccharide

backbone with lipid side chains protruding. It has an approximate molecular weight of

980,000 and is primarily used as a degreasing agent.

Ethanol

The ethanol used to clean the Ottawa sand was purchased from Fisher Scientific,

catalog number A407P-4. For every 100 gallons of ethyl alcohol, the solvent contained

1-gal of ethyl acetate, 1-gal of methyl iso-butyl ketone, and 1-gal of aviation gasoline.

Page 78: Examination of the Effects of Biosurfactant Concentration

-64-

Natural Gas

The natural gas used for hydrate generation was purchased from NexAir, Inc., of

Memphis, TN. The gas has a chemical make up of methane, ethane, and propane in 90,

6, and 4 molar percentages, respectively.

Ottawa Sand

Purified silica sand, i.e. Ottawa sand, was purchased from Spectrum Chemicals of

Gardena, CA. The Ottawa sand has a chemical structure of SiO2 and a purity of 99.0 –

99.9% by weight. The Ottawa sand may also contain trace amounts of such impurities as

titanium oxide, iron oxide, and aluminum oxide. Ottawa sand has a CAS number of

14808-60-7.

Bentonite Clay

Purified sodium bentonite clay, sometimes also known as montmorillonite, was

purchased from Sigma Chemical Company. The bentonite clay has a molecular structure

of Al2O3·4SiO2·H2O and CAS number of 1302-78-9.

Kaolinite Clay

Kaolinite clay, also known as kaolin or china powder, was purchased from

Spectrum Chemical. Kaolin has a chemical structure of Al2O3·SiO2·2H2O, a CAS

number of 1332-58-7, and a face-centered cubic arrangement.

Nontronite Clay

Nontronite clay was purchased from Ward’s Natural Science of Rochester, NY.

Nontronite is very similar to bentonite clay in structure with the aluminum in the

Page 79: Examination of the Effects of Biosurfactant Concentration

-65-

octahedral sheet replaced primarily with iron. Nontronite clay has a principal chemical

structure of Fe2O3·4SiO2·H2O.

Aragonite

Mineral aragonite was purchased from Ward’s Natural Science. Aragonite is a

polymorph of calcite, commonly referred to as calcium carbonate. Aragonite has an

orthorhombic symmetry.

Page 80: Examination of the Effects of Biosurfactant Concentration

-66-

CHAPTER V

RESULTS & DISCUSSION

Scope of Results

Presented in this paper for the first time are determinations of the effect of

biosurfactants at varied concentrations on gas hydrate formation in the presence of

diverse ocean minerals. Biosurfactants interact with different minerals depending on

relative structures. Plots of biosurfactant concentration versus hydrate formation rate and

digital photographs of gas hydrate association are presented which infer adsorptive

interactions between biosurfactant molecules and mineral particles. General observations

are made here about the effect that biosurfactant/mineral interactions have on induction

time of natural gas hydrates and about the systematic ordering of gas hydrates in this

complex scheme. A mechanism is also presented here for hydrate nucleation and

formation in the presence of biosurfactants and porous media. Data are reported using

distilled water instead of seawater in most cases.

Natural Gas Hydrate Formation Rate

The experiments were conducted under constant heat transfer rates and non-

isobaric conditions. With few exceptions, a set of duplicate experiments was run for

Page 81: Examination of the Effects of Biosurfactant Concentration

-67-

nearly all scenarios. In cases of anomalies, triplicates or quadruplicates were run. The

experimental matrix for this investigation is shown in Table 5.1.

Table 5.1. Experimental Matrix

Rhamnolipid+ Concentration OS* OS/Bent. OS/Kaolin OS/Nont. OS/Arag.

0 ppm 3 2 2 2 2 0 ppm (w/seawater) 2 -- -- -- --

10 ppm 3 2 2 -- -- 100 ppm 2 3 2 -- -- 500 ppm 2 2 3 -- -- 1000 ppm 3 2 4 2 2

1000 ppm (w/seawater) 3 -- -- -- --

Emulsan+ Concentration OS OS/Bent. OS/Kaolin OS/Nont. OS/Arag.

10 ppm 3 2 2 -- -- 100 ppm 4 2 1 -- -- 500 ppm 3 2 -- -- -- 1000 ppm 2 3 2 -- --

* Note: OS refers to Ottawa Sand + Note: Unless otherwise noted, all runs were prepared with distilled water.

A measurable property must be chosen or defined on which comparisons can be

based to determine effects of the variables on hydrate formation. A peak formation rate

was chosen as the most accurate and repeatable outcome that could be measured and

calculated. Formation rates were inferred from the decrease in number of moles of gas

from the free gas phase upon hydrate formation. The change in number of moles was

calculated from measured temperatures and pressures recorded at 30 second intervals by

Equation 5.1.

Page 82: Examination of the Effects of Biosurfactant Concentration

-68-

−⋅=∆

ii

i

ff

f

TzP

TzP

RVn (5.1)

In Eq. 5.1, n is the number of moles, P is the pressure of the system, V is the volume of

the pressure vessel, z is the compressibility factor, T is the temperature of the gas, and R

is the universal gas constant. The subscripts i and f refer to initial and final conditions,

respectively. The compressibility factor z was calculated from temperature and pressure

data via the Peng-Robinson cubic equation of state. A sample Peng-Robinson calculation

algorithm is presented in Appendix B for clarity. The formation rate between successive

data points was then calculated from Equation 5.2.

tn

ttnn

rif

ifformation ∆

∆=

−=− (5.2)

Again, n is the number of moles of free gas and t is time. The negative sign in Equation

5.2 denotes that each mole of gas that is consumed from the free gas phase appears as a

mole of gas in the hydrate phase (1:1 molar ratio.)

The peak formation rate was defined as the maximum formation rate value

calculated between two successive data points and is indicated in Figure 5.1. The

overwhelming majority of the data, 81%, exhibited standard deviation values below 0.30

mmol/min and, in most cases (72%), below 0.20 mmol/min. Rare anomalies appeared

with standard deviations as high as 0.66 mmol/min.

Page 83: Examination of the Effects of Biosurfactant Concentration

-69-

1.9891

-2.0000

-1.0000

0.0000

1.0000

2.0000

3.0000

100.00 150.00 200.00 250.00 300.00

Elapsed Time

Rea

ctio

n R

ate

(mm

ol/m

in)

274.00

275.00

276.00

Tem

pera

ture

(K)

Form. Rate Temp.

Figure 5.1. Definition of Peak Formation Rate The basic shape of the rate curve in Fig. 5.1 is consistent from run to run and can

be explained from a thermodynamic standpoint. After nucleation occurs and critical

nuclei of hydrates form, gas hydrates form rapidly and liberate a significant amount of

energy to the bulk gas because of their exothermic heat of formation. Simultaneously, a

drop in pressure occurs rapidly as large amounts of gas occlude into the solid hydrate

phase from the gas phase in a short period of time. These combined effects, a rise in

temperature and a lowering of pressure, signify a rapid increase in the rate of hydrate

formation but also shift the system towards the water-hydrate-gas equilibrium line on a

pressure versus temperature plot. The swing towards equilibrium ultimately prevails and

Peak Formation Rate Value

Page 84: Examination of the Effects of Biosurfactant Concentration

-70-

slows hydrate formation rate (see Eq. 3.4) until the system equilibrates. The point at

which the temperature shift toward the equilibrium line (retardation of formation rate)

overcomes the rapid crystallization defines the peak formation rate of natural gas

hydrates.

A standard error was calculated to be approximately ±0.20 mmol/min for

formation rate variation between successive data points of Fig. 5.1. Therefore, the

standard deviation for peak formation rate in 72% of the data points was within the error

of the rate of formation calculation.

Effect of Biosurfactant Concentration on Formation Rate

The extent of retardation or promotion of natural gas hydrates has not been

studied as a function of biosurfactant concentration. Being anionic and polyanionic,

rhamnolipid and Emulsan expectedly promote hydrate formation in most instances, but

only at sufficient concentrations and only with certain biosurfactant/porous media

combinations.

The critical micellar concentration for micelle-forming surfactants has frequently

been viewed as a measure of a surfactant’s activity. The CMC values of surfactants have

been inferred at hydrate conditions from induction time data [26, p. 74; 30, p. 4178; 31,

pp. 70 – 74]. While these studies make the generality that anionic surfactants increase

hydrate formation rate considerably, no data currently exist for the identification of

biosurfactant concentration effects on hydrate formation in porous media.

Page 85: Examination of the Effects of Biosurfactant Concentration

-71-

Rhamnolipid Concentration Effects on Formation Rate

Figure 5.2 shows the effect of rhamnolipid concentration on gas hydrate

formation in the presence of Ottawa sand, an Ottawa sand/bentonite mixture, and an

Ottawa sand/kaolinite mixture. The error bars represent the standard deviation between

duplicate or triplicate runs, and in some cases, may be within the size of the data symbol.

Smoothed connecting lines have been inserted for easier visualization.

0.00

0.50

1.00

1.50

2.00

2.50

0 200 400 600 800 1000 1200

Concentration (ppm)

Peak

For

mat

ion

Rat

e (m

mol

/min

)

Figure 5.2. Effect of Rhamnolipid Concentration on Gas Hydrate Formation Rate

ο - Ottawa Sand - Ottawa Sand/Bentonite - Ottawa Sand/Kaolinite

Page 86: Examination of the Effects of Biosurfactant Concentration

-72-

Figure 5.2 shows that while rhamnolipid dramatically increases the rate of hydrate

formation in all porous media tested, the biosurfactant concentration at which this

increase occurs is significantly different for each surface.

Pure Ottawa sand and Ottawa sand/bentonite showed immediate effects in the

presence of rhamnolipid at concentrations of less than 100 ppm. Ottawa sand/bentonite

exhibited the most dramatic increase, which is attributable to the nature of bentonite

discussed later. Kaolinite, on the other hand, demonstrated a delay in increase of hydrate

formation rate due to rhamnolipid concentration; only showing significant increases at

concentrations between 100 – 500 ppm.

It should also be noted that, while all porous media exhibited increases in hydrate

formation rate at increased concentrations of rhamnolipid, all sediments attained the same

approximate formation rate at concentrations of 1000 ppm. This fact is supported by

tests with other types of sediment such as aragonite and nontronite. The leveling effect of

gas hydrate formation rate is presented in Figure 5.3.

Each curve in Fig. 5.2 undergoes a transition similar to the critical micellar

transition encountered when plotting surface tension versus concentration in bulk fluids.

The points at which these transitions occur in Fig. 5.2 are markedly different. While

bentonite shows this transition immediately (<100 ppm), Ottawa sand did not exhibit a

transition until the 100 ppm – 500 ppm range. Kaolinite showed the highest

concentration before exhibiting this transition (>500 ppm). This transition may be

indicative of the effective CMC value in solutions with interactions between

biosurfactant and porous media surfaces.

Page 87: Examination of the Effects of Biosurfactant Concentration

-73-

0.00

0.50

1.00

1.50

2.00

Ottawa Sand OS/Bentonite OS/Kaolinite OS/Nontronite OS/Aragonite

Sediment Surface

Form

atio

n R

ate

(mm

ol/m

in)

Figure 5.3. Hydrate Formation Rate at 1000 ppm Rhamnolipid The trend of gas hydrate formation rate converging to a value of ~1.55 mmol/min

at 1000 ppm of rhamnolipid suggests the presence of a rate-controlling process that

dominates at high concentrations. This process assuredly relates to the attainment of a

micellar concentration of rhamnolipid, the adsorption of rhamnolipid on sediment

surfaces, or most likely a combination of both mechanisms.

Emulsan Concentration Effects on Formation Rate

In the presence of porous media, Emulsan presents a much different effect on gas

hydrate formation as a function of concentration. Emulsan concentration and hydrate

Page 88: Examination of the Effects of Biosurfactant Concentration

-74-

formation rate seem to be related in a nearly linear fashion in the presence of Ottawa

sand, bentonite, and kaolinite. Figure 5.4 displays these apparent trends on sediment

surfaces.

0.00

1.00

2.00

3.00

0 200 400 600 800 1000 1200

Concentration (ppm)

Peak

For

mat

ion

Rat

e (m

mol

/min

)

Figure 5.4. Effect of Emulsan Concentration on Gas Hydrate Formation Rate Figure 5.4 shows that Ottawa sand and Ottawa sand/Bentonite systems exhibit

approximately the same general relationship, a linear increase of hydrate formation with

increasing Emulsan concentration. Note that the slope of both lines is approximately the

same. Since Emulsan is a non-micellar surfactant, this phenomenon may be explained by

Bentonite

Ottawa Sand

Kaolinite

Page 89: Examination of the Effects of Biosurfactant Concentration

-75-

the increase of nucleating particles (i.e. Emulsan particles) as the concentration increases.

Bentonite again exhibits a higher rate of formation at each point relative to the other

sediment surfaces. This trend will be discussed further in sections to follow.

Unlike bentonite and Ottawa sand, kaolinite causes no apparent change in hydrate

formation rate upon subsequent addition of Emulsan. These data suggest that in some

fashion kaolinite is deactivating the Emulsan or negating its effect on the hydrate kinetics

and thermodynamics. Furthermore, to accomplish this, kaolinite must effectively

eliminate the increasing number of nucleating sites that Emulsan would present. It is

possible that kaolinite is causing a conformational change of Emulsan whereby the

Emulsan molecule would not resemble the typical size or shape of a hydrate-nucleating

particle.

Several possibilities for this exist. (1) One method by which this could occur

would be Emulsan spreading on the surface of kaolinite that is known to occur on the

surface of oil-water emulsifications [68]. If this spreading were to occur with the

lipophilic groups directed towards kaolinite, Emulsan could not effectively transport gas

to the solution. (2) Emulsan-kaolinite ion exchange is preventing the molecule from

unfurling on the water-gas interface and collecting gas essential for nucleation. (3)

Emulsan-kaolinite ion exchange is preventing precipitation of Emulsan such that small

nucleating particles are no longer present. However, the last two explanations seem less

likely considering kaolinite has a high likelihood of anion repulsion to the polyanionic

Emulsan molecule.

Page 90: Examination of the Effects of Biosurfactant Concentration

-76-

Effect of Porous Media on Formation Rate

Porous media has been documented as having a significant surface effect on the

rate of gas hydrate formation and dissociation in the absence of any surfactant or

biosurfactant. Ginsburg, et al., have presented the preference of gas hydrates toward

areas of high porosity [40, p. 237]. Uchida, et al., also noted that the dissociation

temperature of methane hydrates varied in an inversely proportional fashion to pore

diameter in porous glass of pore diameter between 100 and 500Å [43, p. 3659]. Cha, et

al., proved that sodium montmorillonite promoted hydrate formation due to a “possible

surface ordering” effect in which the hydroxide edges of montmorillonite act as one side

of the hydrate lattice [41, p. 6492].

Furthermore, peak formation rate data of this study also show a trend toward

specific surfaces under the influence and absence of biosurfactants. Figure 5.5 shows the

peak formation rate of natural gas hydrates in distilled water and on various mineral

surfaces.

Page 91: Examination of the Effects of Biosurfactant Concentration

-77-

0.540.45

0.75

0.90

0.40

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

1

Sediment Surface

Form

atio

n R

ate

(mm

ol/m

in)

Ottawa Sand Bentonite Kaolinite Nontronite Aragonite

Note: No Biosurfactant Present Figure 5.5. Effect of Sediment on Peak Formation Rate in Distilled Water Figure 5.5 shows that Ottawa sand, kaolinite, and nontronite all have base values

that approach 0.45 mmol/min. A near doubling of formation rate occured when bentonite

is present as predicted by Cha, et al. Interestingly aragonite induced an increase in

hydrate formation rate to a value of 0.75 mmol/min.

Cha, et al., hypothesized that the hydroxyl edges of the bentonite clay were

forming hydrogen-bonded links between the gas hydrate lattice, incorporating itself into

the hydrate structure [41, p. 6494]. If this hypothesis were completely true, then it would

be expected to see a similar effect with nontronite which has a similar layer structure to

Page 92: Examination of the Effects of Biosurfactant Concentration

-78-

bentonite. This was not the case. As a matter of fact, there appeared to be a depression

in formation rate between nontronite clay and the control surface, Ottawa sand.

Alternatively, the increase in hydrate formation rate caused by bentonite is

apparently due to the intercalation of water molecules and structuring in the water phase

between the basal planes. Kaolinite has no noticeable intercalation of water due to

strong, well-structured hydrogen bonding between the basal planes of its platelets [47, p.

83]. Nontronite also has no appreciable intercalation of water as evidenced by its relative

lack of swelling in water media. However, bentonite intercalates not only water

molecules but also surfactant molecules quite easily [50, p. 367]. These data suggest that

the key to hydrate promotion by clay surfaces may not be in surface interactions, but

rather in interlayer interactions.

Aragonite presents a different type of anomaly. Aragonite has no layers to

intercalate with water, yet it induced a significant increase in natural gas hydrate

formation rate with distilled water. One possible explanation is that the CaCO3 structure

organizes the water layer very near the water-gas interface. Aragonite is a microporous

structure with ample surface area to structure water and promote hydrate growth.

Support for this explanation is presented in depth in a following section.

Adsorption of Biosurfactants on Porous Media

Quantitative adsorption tests were conducted to determine the relative extent of

biosurfactant adsorption by each type of sediment surface. From changes in surface

tension of entering and exiting water solutions, the adsorptivity of each surfactant was

measured. The preference of each surfactant for a particular surface was then inferred by

Page 93: Examination of the Effects of Biosurfactant Concentration

-79-

determining surfactant concentration from the surface tension versus concentration

calibration curve (See Appendix A for curves, see Chapter IV for procedure.) The

surface tension differences of these tests are shown below in Table 5.2.

Table 5.2. Biosurfactant Selective Adsorption Test Rhamnolipid on Ottawa Sand Emulsan on Ottawa Sand Concentration Standard Surf. Tension ∆(S.T.) Standard Surf. Tension ∆(S.T.)

(ppm) (mN/m) (mN/m) (mN/m) (mN/m) (mN/m) (mN/m)10 4.7 2.4 -2.3 -3.2 -2 1.2

100 4.5 6.4 1.9 -0.7 -0.4 0.3 1000 -0.1 -0.2 -0.1 -0.8 -0.6 0.2

Rhamnolipid on Bentonite Emulsan on Bentonite

10 4.7 13.3 8.6 -3.2 10 13.2 100 4.5 21.3 16.8 -0.7 10.4 11.1

1000 -0.1 9.7 9.8 -0.8 4.7 5.5 Rhamnolipid on Kaolinite Emulsan on Kaolinite

10 4.7 0.7 -4 -3.2 -3.7 -0.5 100 4.5 10.3 5.8 -0.7 0 0.7

1000 -0.1 1.3 1.4 -0.8 0.5 1.3

Table 5.2 shows that, for the case of rhamnolipid, there was notable adsorption by

the blank trial at lower concentrations (<100 ppm). This trend was expected, as an

anionic surfactant is susceptible to being adsorbed onto metal surfaces such as the wire

mesh used to keep sediment from being eluted through the column.

For similar reasons, the reverse trend for Emulsan is curious. The fact that no

negative adsorption is perceived at these concentrations may be attributable to the

valence structure of Emulsan. The polyanionic biosurfactant may be repulsed notably by

the PVC walls as the walls are subsequently hydrated. Repulsion followed by hydration

of the PVC walls would both concentrate Emulsan in the bulk and concentrate water

Page 94: Examination of the Effects of Biosurfactant Concentration

-80-

away from the bulk. These combined effects would give Emulsan a higher concentration

after being passed through the column than before and would result in a decrease in

surface tension instead of the expected rise. This observation is present in other porous

media tests also.

In the presence of rhamnolipid, bentonite clay seemed to have the largest effect on

adsorption. This fact was predicted by the presence of the hydroxyl sites along the edges

of a bentonite platelet that easily serve to adhere an anionic surfactant. Also the

intercalation of rhamnolipid into bentonite interlayers surely plays a significant role.

While each surface shows some adsorption at 100-ppm concentrations, only bentonite

shows adsorption at 10-ppm concentrations. This fact is very telling, suggesting that a

weak adsorption is occurring at 100 ppm which is absent at 10 ppm for all surfaces

except bentonite. This absence can be accounted for by the lack of micelles at 10-ppm

concentrations. In other words, in the presence of sand or sand and kaolinite,

rhamnolipid is adsorbed in clusters of micelles, not as individual molecules.

The same seems to be true at 1000-ppm concentration but is easily explained. At

1000 ppm, a great amount of adsorption would have to take place to noticeably change

the surface tension. A 1000-ppm concentration is well into the horizontal portion of the

surface tension curve. This explanation is also supported by a similar trend in the

Emulsan test. This explanation attributes to the large adsorptive capacity of bentonite

since a large amount of rhamnolipid must be adsorbed to noticeably change surface

tension in this region.

Emulsan shows little or no adsorption on sand or a sand/kaolinite mixture at any

concentration. Since no micelles form with Emulsan, there is no micelle clustering at the

Page 95: Examination of the Effects of Biosurfactant Concentration

-81-

surface of the sediment as was suggested in the previous explanation. However, as

before, bentonite showed a relatively high affinity for the Emulsan molecule at all

concentrations. Presumably, this fact is caused by the attraction of the polyanionic

molecules to the net positive hydroxyl edges of the bentonite platelet and possibly also in

the interlayers of bentonite.

Adsorption and Biosurfactant Concentration Related to Formation

Four adsorption isotherms commonly appear in adsorption or chemisorption

kinetics: the L-type isotherm, the S-type isotherm, the C-type isotherm, and the H-type

isotherm [69, p. B-277]. The shape of each curve, discussed in Chapter III, indicates the

type and extent of adsorption or chemisorption by an ion or molecule to a solid surface.

The shape of each biosurfactant concentration versus gas hydrate formation rate

plot has a very distinctive curve which is porous media dependent. Sloan has also

proposed that gas enclathration is analogous to adsorption and desorption [3, pp. 208 –

211]. The current data suggest that the shape of the concentration versus hydrate

formation curve is not accidental, but likely dictated by the adsorptive characteristics of

each mineral-biosurfactant interaction. Thus, the affinity of a biosurfactant molecule can

be qualitatively deduced from the curvature of these formation plots. In the case of

rhamnolipid, a necessary assumption is that at higher concentrations of rhamnolipid,

adsorption and desorption are in equilibrium and free micellar shapes dictate the

formation rate regardless of porous media present.

Ottawa sand in the presence of rhamnolipid had a relatively intermediate slope at

lower concentrations (<300 ppm) and effectively leveled off to a stable peak formation

Page 96: Examination of the Effects of Biosurfactant Concentration

-82-

rate value of 1.64 mmol/min. Figure 5.6 shows the change in peak hydrate formation rate

due to rhamnolipid concentration in an Ottawa sand pack. Figure 5.6 and Figures 5.7 and

5.8 to follow were fitted with cubic splines to better visualize the overall trend in gas

hydrate formation.

0.00

0.50

1.00

1.50

2.00

0.00 200.00 400.00 600.00 800.00 1000.00 1200.00

Concentration (ppm)

Peak

For

mat

ion

Rat

e (m

mol

/min

)

Figure 5.6. Rhamnolipid Concentration Vs Peak Formation Rate in Ottawa Sand The plot of Fig. 5.6 is indicative of a typical L-type (Langmuir) plot, which rises quickly

but levels off to a constant quantity. The L-type curve for Ottawa sand and rhamnolipid

suggests that adsorbate-adsorbent interactions are relatively high and may be indicative

of chemisorption. The means by which this occurs is unclear, however, since the surface

of Ottawa sand should be a net negative surface and rhamnolipid is an anionic surfactant.

The chemisorption may be an interaction between micelles adsorbed on the surface of

Page 97: Examination of the Effects of Biosurfactant Concentration

-83-

sand (admicelles). This fact would explain the lack of appreciable change in peak

formation rate in the 10-ppm concentration range before micelles form.

In the presence of bentonite, the rhamnolipid concentration versus peak formation

rate curve takes on a similar shape with a much steeper initial slope. This initial slope is

followed by a maximum and then a slight decline back to a micelle-dictated value of 1.50

mmol/min. This curve is presented in Figure 5.7.

0.00

1.00

2.00

3.00

0.00 200.00 400.00 600.00 800.00 1000.00 1200.00

Concentration (ppm)

Peak

For

mat

ion

Rat

e (m

mol

/min

)

Figure 5.7. Rhamnolipid Concentration Vs Peak Formation Rate in Bentonite Clay The steepness of the rise in the lower concentration range for bentonite suggests a much

stronger affinity of bentonite for the rhamnolipid molecule than exhibited by Ottawa

sand. The rise is so extreme that it borders on an H-type isotherm, indicating very strong

adsorbate-adsorbent interaction.

Page 98: Examination of the Effects of Biosurfactant Concentration

-84-

The presence of a peak value and decline upon subsequent addition of

rhamnolipid is a curious one. This peak may be brought about by a filling of the

interlayer of bentonite due to intercalation of rhamnolipid. Once this interlayer is filled

by admicelles, the interlayer is deactivated, shifting the formation back to a free-micelle

directed formation rate. This explanation is under the assumption that adsorption of

rhamnolipid is notable in the interlayers in addition to adsorption along the hydroxylated

edges of the bentonite platelets.

Kaolinite in the presence of rhamnolipid gives a very distinctive S-shaped

adsorption curve. The S-shaped adsorption curve results when adsorbate-adsorbate

interactions are stronger than the adsorbate-adsorbent interactions causing clustering of

adsorbate molecules near the surface. This clustering of adsorbate molecules for the case

of rhamnolipid is micelle formation. The relative absence of adsorbate-adsorbent

interaction in the case of kaolinite may be the result of anion repulsion. The S-curve for

kaolinite is presented in Figure 5.8.

Page 99: Examination of the Effects of Biosurfactant Concentration

-85-

Effect of Rhamnolipid on Kaolinite

0.00

0.50

1.00

1.50

2.00

0 200 400 600 800 1000 1200

Concentration (ppm)

Peak

For

mat

ion

Rat

e (m

mol

/min

)

Figure 5.8. Rhamnolipid Concentration Vs Peak Formation Rate in Kaolinite Clay Through weak adsorbate-adsorbent interactions, kaolinite may deactivate a

number of the rhamnolipid molecules not allowing them to form micelles or admicelles.

These interactions could result in a significant increase in the CMC value of rhamnolipid

and could be responsible for the slow formation rates at low rhamnolipid concentrations

with respect to Ottawa sand or bentonite.

Emulsan in the presence of porous media presents a very different curve of

concentration versus peak formation rate (see Fig. 5.4.) With each porous media tested,

this curve is approximately linear, indicative of C-type adsorption. Equations 5.3, 5.4,

and 5.5 show the best fit curves for peak hydrate formation rate versus Emulsan

concentration in the presence of bentonite, Ottawa sand, and kaolinite, respectively.

Page 100: Examination of the Effects of Biosurfactant Concentration

-86-

9893.00011.0 +⋅= EMPF Cr (5.3)

6334.00008.0 +⋅= EMPF Cr (5.4)

43.0=PFr (5.5)

In the above equations, rPF is the peak hydrate formation rate measured in mmol/min, and

CEM is the concentration of Emulsan measured in ppm.

The C-type adsorption is characterized by a constant relative affinity of adsorbate

to adsorbent. If this is the case, then bentonite clay shows the highest affinity for

Emulsan of any media while kaolinite clay shows no affinity at all. Curiously, the slope

of the C-type straight line for the Ottawa sand system and the Ottawa sand/bentonite

system are nearly the same (see Figure 5.4.) This fact suggests that the relative affinity of

Emulsan to Ottawa sand and bentonite is approximately the same. The shift in intercept

of these two lines can be explained by the additional structuring effect that bentonite

possesses over Ottawa sand as discussed previously.

The C-type adsorption curve for kaolinite in the presence of Emulsan has a slope

of zero. It is, however, unlikely that Emulsan has no affinity for kaolinite. A more

plausible explanation would be that the lipophilic groups of the Emulsan are adhering to

the surface of the kaolinite clay, much like a commercial surfactant would adhere to an

organic particle. The adhesion of the lipophilic groups to the kaolinite would deactive

those moieties from solubilizing gas but more importantly would destroy the tertiary

structure of Emulsan. The unfurling of the Emulsan molecule could prohibit it from

providing nucleation sites for the gas hydrates to originate.

Page 101: Examination of the Effects of Biosurfactant Concentration

-87-

Induction Time

Induction time is defined for the subject experiments as the difference between

the onset of hydrate crystallization and the time at which the gas hydrate equilibrium line

is crossed (point of supersaturation). Induction time gives a simple understanding of the

time required to accumulate a critical nucleus of hydrate crystals upon supersaturation

[73]. Studies recently have hinted at a possible correlation between induction time and

the critical micellar concentration of rhamnolipid [26, p. 74; 30, p. 4178; 42, p. 977]. The

length of induction time is shown below in a representative temporal plot of pressure

versus temperature noting the amount of supersaturation.

240.00

250.00

260.00

270.00

280.00

290.00

300.00

310.00

320.00

270.00 275.00 280.00 285.00 290.00 295.00

Temperature (K)

Pres

sure

(PSI

G)

Run #100 Formation Equil. Run #100 Decomposition

Ottawa SandDistilled Water/Rhamno (1000 ppm)Moisture (sat'd)

Ti

Figure 5.9. Induction Time in Gas Hydrate Growth

Induction Time

Page 102: Examination of the Effects of Biosurfactant Concentration

-88-

Each run was given a maximum of 48 hrs to begin forming hydrates, and

induction time was calculated for each case in which hydrates formed. A select few

experiments had induction times that could not be determined due to slow hydrate

formation. Eleven cases did not form hydrates to any appreciable extent or had

indeterminate induction times.

For cases where hydrates formed, induction times ranged from less than one hour

to as long as 34 hours. To demonstrate that hydrates would eventually form if given

ample time, one experiment was left running for longer than the allotted 48 hour

maximum and had an induction time of over 75 hours. Table 5.3 below shows the

average induction time for all runs that formed hydrates and the standard deviation. In

Table 5.3, OS denotes Ottawa sand, BE denotes bentonite clay, KA denotes kaolinite

clay, NO denotes nontronite clay, and AR denotes aragonite.

Page 103: Examination of the Effects of Biosurfactant Concentration

-89-

Table 5.3. Induction Time

Rhamnolipid

Conc. (ppm)

Ti, OS (hrs)

Stand. Dev.

Ti, OS/BE (hrs)

Stand. Dev.

Ti, OS/KA (hrs)

Stand. Dev.

0 7.77 5.47 2.32 1.93 3.73 2.19 10 6.67 -- 6.43 2.20 3.90 2.20

100 1.50 0.42 2.31 1.45 3.23 2.03 500 40.63 49.13 1.63 0.68 4.94 2.59

1000 9.04 5.70 14.77 15.59 1.94 0.52

Emulsan

Conc. (ppm)

Ti, OS (hrs)

Stand. Dev.

Ti, OS/BE (hrs)

Stand. Dev.

Ti, OS/KA (hrs)

Stand. Dev.

0 7.77 5.47 2.32 1.93 3.73 2.19 10 6.59 7.91 3.60 3.02 32.23 --

100 1.51 -- 2.23 1.56 10.88 -- 500 10.80 4.03 3.00 1.09 -- --

1000 14.84 11.68 5.17 3.99 3.48 0.71

Rhamnolipid Rhamno/Seawater

Conc. (ppm)

Ti, OS/NO (hrs)

Stand. Dev.

Ti, OS/AR (hrs)

Stand. Dev.

Ti, OS (hrs)

Stand. Dev.

0 9.19 -- 0.60 0.24 5.63 1.26 1000 6.38 5.32 17.32 23.50 3.92 1.24

Table 5.3 shows that there appears to be no correlation between induction time

and biosurfactant concentration. The data presented seem to be erratic and irreproducible

as demonstrated by the high standard deviation values, in some cases exceeding the

average induction time values. This fact may possibly be explained by the difficulty in

reproducing accurately the quantity of nucleating sites when porous media is present.

Other studies that have suggested a correlation between biosurfactant concentration and

induction time have been in the absence of porous media.

Page 104: Examination of the Effects of Biosurfactant Concentration

-90-

Heat and Gas Transfer Effects on Formation Rate

Other researchers have examined the role of heat and mass transfer rates on gas

hydrate formation [44; 71, pp. 1069 – 1071; 74, pp. 301 – 302; 75, p. 465; 76]. While

mass transfer affects the rate at which gas molecules are contacted with prospective

hydrate cages, heat transfer dictates the rate at which the heat of formation is removed,

and thus, sets the location of the operating point relative to the equilibrium curve. With

these facts in mind, the experiment was conducted to minimize deleterious mass transfer

and heat transfer effects.

Figure 5.10 compares experimental runs limited by heat and mass transfer to

experimental runs where heat and mass transfer limitations are negligible. For

comparison purposes, the data have been normalized by dividing each data set by the

fastest peak formation rate within each set. The data represented in Fig. 5.10 were

obtained with 1000 ppm rhamnolipid-distilled water solution on a Ottawa sand/bentonite

pack.

Page 105: Examination of the Effects of Biosurfactant Concentration

-91-

0

0.2

0.4

0.6

0.8

1

0 200 400 600 800 1000 1200

Concentration (ppm)

Peak

For

mat

ion

Rat

e (n

orm

aliz

ed)

Not Heat & Mass Transfer Limited Heat & Mass Transfer Limited

Figure 5.10. Effect of Heat and Mass Transfer Limitation Figure 5.10 shows that under heat and mass transfer limited conditions, an inverse

correlation exists between concentration of rhamnolipid and peak formation rate. In

other words, under the limitations of heat and mass transfer, rhamnolipid is a hindrance in

the formation rate of gas hydrates.

Mass transfer limitations resulted in the experiments when insufficient gas ports

were available for gas access. Possibly, rhamnolipid molecules at the water-gas interface

promoted hydrates to block those gas access ports, thereby providing a barrier rather than

a means of gas transport when interfacial area is limited. Another explanation could be

that the presence of rhamnolipid is altering the bulk heat capacity of the solution such,

that under heat transfer limited conditions, the heat of formation cannot be removed

Page 106: Examination of the Effects of Biosurfactant Concentration

-92-

effectively. In addition, the Teflon container used previously, as opposed to the thinner

polypropylene sample container, allowed for less heat transfer into and out of the sample.

Effect of Electrolytes on Formation Rate

Electrolyte solutions have been extensively documented to inhibit gas hydrate

formation [36, pp. 70 – 73; 38, pp. 22 – 27; 77, p. 1719 – 1721; 78]. This fact has been

theorized to be attributable to the change in colligative properties of the liquid phase

whereby freezing point may be depressed. The inhibition may also have to do with the

energy required to expunge the electrolyte particles into the interstitial water and out of

the gas hydrate lattice.

A series of experiments were undertaken to determine the approximate effect of

Gulf of Mexico seawater on the rate of hydrate formation in the presence and absence of

biosurfactant. The formation rate in seawater was compared to distilled water in an

Ottawa sand pack and to a 1000 ppm distilled water-rhamnolipid solution in an Ottawa

sand pack. The data are presented in Table 5.4.

Table 5.4. Effect of Electrolytes on Gas Hydrate Formation

Peak Formation Rate

(mmol/min) † STDEVPeak Formation Rate

(mmol/min) ‡ STDEV DI Water 0.54 0.04 1.64 0.32 Seawater 0.39 0.04 1.47 0.40 † - No surfactant present ‡ - 1000-ppm rhamnolipid present

Page 107: Examination of the Effects of Biosurfactant Concentration

-93-

As expected, it appears that the peak formation rate is depressed by seawater in

the presence and absence of biosurfactant, but the value differences are very near the

limits of experimental error since the saltwater concentration is low.

Gas Hydrate in Porous Media, Preference Trends

In porous media, gas hydrates have a preference for high porosity [40, p. 237].

This fact is intuitive because gas hydrates, like ice, expand upon formation. Therefore, it

is expected that gas hydrates would form in regions of highest gas-water interfacial area

with the most room to expand. This observation is witnessed as gas hydrates routinely

form around the mouth of the sample cup and around the gas ports into the media pack.

Beyond this intuitive observation, however, is the preference of gas hydrates to

agglomerate on certain surfaces in preference to other surfaces when given the choice.

For instance, in the presence of Ottawa sand as the only porous media, hydrates choose to

cluster about the stainless steel RTD probe that just contacts the surface of the sand. Not

only do gas hydrates prefer the stainless steel to silica, but also gas hydrates often bridge

from the sample cup to the stainless steel walls of the reaction pressure vessel. This

phenomenon is shown in Figure 5.11.

Page 108: Examination of the Effects of Biosurfactant Concentration

-94-

Figure 5.11. Preference of Gas Hydrates to Stainless Steel Over Silica (OS) This preference of gas hydrates to stainless steel was present in all cases where

Ottawa sand or kaolinite was the principle surface of interest. Only in the cases of

bentonite, nontronite, and aragonite did the preference shift.

Quite noticeably gas hydrates prefer, given the opportunity, to crystallize on the

surface of the smectites over the silica or stainless steel. In all cases of bentonite and

nontronite (with rhamnolipid, Emulsan, or neither present), massive hydrates formed

around the mouth of the sample cup around the semi-circular region containing the

smectite. Figure 5.12 clearly demonstrates this fact.

Vacancy created by RTD probe

OS, Rhamno. (1000 ppm)

OS, Emulsan (1000 ppm)

Hydrate bridging to SS vessel wall

Page 109: Examination of the Effects of Biosurfactant Concentration

-95-

Figure 5.12. Preference of Gas Hydrates to Smectites This preference possibly occurs due to structuring effects of the smectite clays on

the water phase versus the other surfaces. It has been proposed that the intercalation of

water and possibly surfactants into the smectite interlayer causes an ordered system and

dually acts as a nucleation site for subsequent hydrate formation [41, p. 6494].

Aragonite was mentioned previously to promote gas hydrate formation in the

presence of distilled water. Under this condition, gas hydrates formed prominently on the

surface of the aragonite as well as the stainless steel RTD probe. However, when

Bentonite

Nontronite

Ottawa Sand

Ottawa Sand

Page 110: Examination of the Effects of Biosurfactant Concentration

-96-

rhamnolipid was introduced, the preference shifted to the shielded RTD probe. This

observation suggests that aragonite effectively organizes water and promotes gas hydrates

as long as distilled water is the only medium. But when an anionic surfactant such as

rhamnolipid is introduced, the surfactant is attracted to the metal surface and promotes

gas hydrates away from the aragonite surface. Figure 5.13 clearly shows this

observation.

Figure 5.13. Preference of Gas Hydrates to Aragonite with No Surfactant Present

Gas Hydrate Packaging, Biosurfactant Ordering

Gas hydrates are known to assume such packing arrangements in ocean sediments

as massive, nodular, dispersed, dendritic, needle-like, and stratified [23, pp. 88 – 102].

No Surfactant Present 1000 ppm Rhamnolipid Present

Aragonite Aragonite

Page 111: Examination of the Effects of Biosurfactant Concentration

-97-

For this investigation, four types of hydrate packing arrangements dominated: massive,

nodular, dendritic, and needle-like. Dendritic and needle-like hydrates were the most

commonly observed forms of hydrates in the absence of biosurfactants or when

biosurfactants were at low (<l00 ppm) concentrations. However, when concentrations of

biosurfactant was greater than or equal to 100 ppm, hydrates took on a general

appearance of massive, rounded hydrates with intermittent cases of nodular hydrates.

These general hydrate conformations are presented in Figure 5.14.

Figure 5.14. Gas Hydrate Packing Arrangements

Nodular Massive

Dendritic Needle-Like

Page 112: Examination of the Effects of Biosurfactant Concentration

-98-

Makogon has attributed the presence of needle-like hydrates to hydrate crystal

growth from water in the vapor phase attaching to an already growing hydrate nucleus

[23, pp. 88 – 89]. Yet the shift from a needle-like morphology to a massive, amorphous

morphology upon increase in biosurfactant concentration suggests that surface tension

has a fundamental role. If needle-like hydrates are being grown from vapor phase water,

then a lowering of surface tension and the capillary effect may be dictating the close-knit

packing arrangement observed when biosurfactant is present at appreciable

concentrations. The surface activity at higher concentrations is dictating the crystal

growth and allowing for a higher ordered structure. This fact is supported by the

apparent transition from primarily needle-like to primarily dendritic to primarily massive

hydrates upon increase in biosurfactant concentration.

In fact, it may be reasoned that the case of dendritic hydrates as an intermediary

has significance. At 10 ppm concentrations, the solution does not have a high enough

concentration to promote hydrates and crystal growth is initiated from the vapor phase.

However, as gas hydrates form, water is removed from the solution leaving behind the

biosurfactant in the interstitial water. As some point, this effect could concentrate the

interstitial water to a point beyond where surface tension plays a role. At this point,

hydrates would grow to a more uniform, rounded massive structure. This explanation is

supported by Fig. 5.14 where both massive structures and dendritic crystals can be seen.

Also notably, aragonite with distilled water exhibits dendritic hydrates as

compared to needle-like hydrates (See Fig. 5.13.) This observation supports the

explanation that aragonite inherently structures the water near the water-gas interface

such that a more ordered system appears rather than a random needle-like structure.

Page 113: Examination of the Effects of Biosurfactant Concentration

-99-

Dispersed Sediment in Massive Hydrates

One phenomenon that is significant and should not be overlooked is the

dispersion of fine sediment particles within the gas hydrate matrix. The most noticeable

case of this event was with nontronite clay. When hydrates form massive collections on

the surface of the nontronite semi-circle, the massive hydrate mound takes on a faint lime

green appearance. This pale, yellowish-green tint is due to very small particles of

nontronite, a yellowish-green clay, being dispersed throughout the gas hydrate lattice

structure. The evidence is presented in Figure 5.15.

Figure 5.15. Nontronite Dispersed Within Hydrate Matrix This occurrence is not limited to nontronite, however. The phenomena also

persists with kaolinite and bentonite but is not apparent until the hydrate mass has

decomposed, leaving behind a puddle of white-colored water or brownish colored water.

Faint Lime Green Color Within Gas Hydrates

Page 114: Examination of the Effects of Biosurfactant Concentration

-100-

Two plausible explanations for this are offered. The first explanation is that small

hydrates crystals adhere to fine grain particles and are transported through water-hydrate

capillaries carrying the fine grain particles with them. The second explanation is that

hydrates grow around the fine particles building up layers which effectively displace the

fine grain particles from their original location on the surface of the cup. The first

explanation would be relatively energy intensive compared to the second possibility.

However, biosurfactants’ lowering of surface tension would lessen the energy needed to

carry a fine grain particle through a hydrate capillary making it easier for hydrates to

form rapidly.

Page 115: Examination of the Effects of Biosurfactant Concentration

CHAPTER VI

CONCLUSIONS

This thesis is the first to document the extent of biosurfactant catalyzation of

natural gas hydrates as a function of concentration and to classify their interactions with

ocean-type sediments.

Significant results were obtained by testing natural gas hydrate formation in the

presence of two biosurfactants, rhamnolipid and Emulsan, and diverse porous media.

The results were dramatically different depending on which biosurfactant was chosen,

hinting at some underlying micellar effect and interaction of media surface with

biosurfactant.

Adsorption

Bentonite was the only surface to effectively adsorb both rhamnolipid and

Emulsan. Its positively charged hydroxyl edges and interlayers effectively attract the

anionic surfactants. Bentonite also has the ability to trap water and biosurfactant within

its interlayers through intercalation. This adsorptive ability along with bentonite’s ability

to structure water give the clay an advantage in hydrate formation that is accentuated in

the presence of biosurfactant.

Page 116: Examination of the Effects of Biosurfactant Concentration

-102-

Each plot of biosurfactant concentration versus peak formation rate, regardless of

porous media, exhibits the shape of a type of adsorption curve. While hydrate formation

has been likened before to the adsorption process, these data suggest that the type of

adsorption that biosurfactants exhibit on clay surfaces may dictate their ability to catalyze

gas hydrates. Bentonite does this most effectively and therefore exhibits the greatest

catalyzation of natural gas hydrates at each concentration, regardless of biosurfactant or

porous media used.

While differing at lower concentrations (<200 ppm), sand and kaolinite exhibit

similar curves, suggesting that the adsorption of biosurfactants molecules on these

surfaces is minimal (or not present) relative to adsorption exhibited by bentonite.

Adsorption tests also showed that the adsorption of rhamnolipid on sand and kaolinite

might be attributable to admicelles and not to the adsorption of individual biosurfactant

molecules.

Other experiments showed that adsorption of Emulsan onto sediment surfaces is

likely through a C-type or constant affinity attraction (independent of concentration.)

However, adsorption tests ran with Emulsan showed that only bentonite adsorbed

Emulsan to an appreciable extent. The C-type slope of Emulsan concentration versus

peak formation curve could be a manifestation of an increase of nucleating sites as

concentration increases. Yet this explanation does not account for the zero slope of the

curve in the presence of kaolinite. This deactivation of Emulsan on kaolinite may be

attributable to Emulsan spreading on the surface of kaolinite and no longer resembling

the conformation required to promote hydrate catalyzation.

Page 117: Examination of the Effects of Biosurfactant Concentration

-103-

Formation Rate

Along with the previously mentioned adsorption related trends, trends of peak

formation rate occur for individual sediments. When tested with distilled water,

bentonite again demonstrated the most proficiency in forming hydrates. Nontronite, a

similar smectite clay, did not show any proficiency at catalyzing hydrates as expected.

Apparently, the intercalation of water by bentonite is fundamental in its ability to

kinetically promote gas hydrate formation. To some degree, this fact also disproves the

notion that the hydroxyl edges of the clay are the key players in water structuring for gas

hydrate promotion. The intercalation of water must play another more fundamental role.

Aragonite shows great proficiency for forming gas hydrates with distilled water.

The aragonite structure must be structuring the water in a manner that is kinetically

favorable in the sense of hydrate formation, yet no hydrates visibly form on the surface of

the aragonite. This anomaly is curious but was not studied in depth. All other surfaces

showed no catalytic effect on hydrate formation.

At high concentrations of rhamnolipid (1000 ppm), the peak formation rate

converged for all porous media. Again, a micellar action is suggested. Possibly this

concentration is adequate to negate all adsorption effects (adsorption is at a steady-state

maximum) and allow for hydrate catalysis in the free micellar state. That is, at this point,

free micelles dictate the rate of formation, not adsorbed micelles or adsorbed rhamnolipid

molecules.

Several experiments were also run to determine the effect of heat and mass

transfer on hydrate formation under the conditions of biosurfactant and porous media

interaction. When heat and mass transfer limit hydrate formation, many of the above

Page 118: Examination of the Effects of Biosurfactant Concentration

-104-

trends are reversed. Most notably, reversal of rhamnolipid concentration versus peak

formation rate occurs in the presence of bentonite. For this scenario, hydrates form very

rapidly in distilled water, but formation slows dramatically when rhamnolipid availability

approaches 1000 ppm.

Hydrate Induction

A correlation between biosurfactants and induction does not exist for the packed

media of these tests. Perhaps this fact is a result of an improper cleaning of tiny particles

from the sediment serving as nucleation sites.

Structure and Preference

The presence of biosurfactant has been shown to have a remarkable effect on the

types of hydrates formed. When hydrates are formed from distilled water, crystal growth

is slow and needle-like in nature. However, when biosurfactant concentration is

increased, the hydrate shape goes through a progression from needle-like, to dendritic, to

nodular, and finally massive. When biosurfactant is in high concentration, the massive

hydrate mounds take on a rounded, structured look accounted for by the ordering

associated with the biosurfactant molecules. Hydrates also seem to grow through

capillary effects where small hydrate clusters are pulled to the surface through hydrate

capillaries. Biosurfactants reduce the energy required to travel through these capillaries.

Natural gas hydrates have a media surface preference. If given the opportunity,

hydrates form at the points of highest porosity and gas concentration. If given the choice

between multiple surfaces, bentonite or nontronite was the preferred choice. Closely

behind the smectites, cold stainless steel was the usual choice. In the presence of any

Page 119: Examination of the Effects of Biosurfactant Concentration

-105-

other type of sediment, hydrates preferred to form near the stainless steel RTD shield

than any other surface.

Scientific Significance

If natural gas hydrates on the ocean floor are ever to become a viable natural

resource, a fundamental understanding of their environment, their formation, and their

decomposition is needed. Natural gas hydrates exist in a symbiotic world where

microbes, porous media, natural oil and gas, and gas hydrates interact. The promotion of

natural gas hydrates by microorganisms and their metabolic functions is shown in this

thesis to be substantial. Conversely, many organisms need the methane that natural gas

hydrates encase.

This thesis helps explain the large amount of gas hydrates that have been

discovered in ocean sediments in recent years. Biosurfactants produced naturally by

microbes in the ocean-floor ecology catalyze gas hydrate formation in that environment.

These gas hydrates store vast quantities of both biogenic and thermogenic hydrocarbon

gases.

Furthermore, better understanding of the method by which gas hydrates form as

determined in this thesis, especially in marine environments, serves four major purposes.

(1) The means of farming natural gas hydrates in the ocean floor can be better

understood. (2) The additional knowledge gained about gas hydrate formation should

assist the oil and gas industry in foreseeing seafloor instabilities. (3) The understanding

of gas hydrate kinetics affects alternative fuels potential and alternative fuels storage. (4)

Results from this study should be helpful in predicting the location of hydrates in ocean

Page 120: Examination of the Effects of Biosurfactant Concentration

-106-

sediment, given the stratigraphy of the formations, and in stabilizing global climate

change. Benefits are also foreseen in predicting massive, dispersed or nodular hydrates

in the sea floor.

Summary

1) Biosurfactants promote hydrate growth by increased solubility of gas, reduced

capillary forces, and possibly structuring of water.

2) Biosurfactant concentration versus peak formation rate plots give curves for

each type of sediment indicative of the sediment’s adsorptive properties.

3) Bentonite interlayers and hydroxyl edges increase gas hydrate formation rate

over other surfaces.

4) Bentonite interlayers and hydroxyl edges adsorb individual rhamnolipid

molecules while sand and kaolinite adsorb micelle structures.

5) Kaolinite deactivates Emulsan, not allowing any increase in peak hydrate

formation rate possibly due to Emulsan spreading on kaolinite surfaces.

6) Gas hydrates form needle-like, nodular, dendritic, and massive structures

depending on the biosurfactant-porous media combination.

7) When presented with a choice, gas hydrates prefer to form on porous media

surfaces in the following order: bentonite or nontronite > aragonite or steel >

sand or kaolinite.

8) If heat and mass transfer are rate-limiting steps in hydrate formation,

biosurfactant/porous media effects may be overshadowed.

Page 121: Examination of the Effects of Biosurfactant Concentration

-107-

REFERENCES [1] M. Faraday, “On Fluid Chlorine,” Transactions of the Royal Society of London,

13 March 1823. [2] E.G. Hammerschmidt, “Formation of Gas Hydrates in Natural Gas Transmission

Lines,” Ind. Eng. Chem., vol. 26, pp. 851 – 855, Aug. 1934. [3] E.D. Sloan, Clathrate Hydrates of Natural Gas. NY: Marcel Dekker, 1990. [4] B.D. Lanoil, et al., “Bacteria and Archaea Physically Associated with Gulf of

Mexico Gas Hydrates,” App. Env. Microbio., vol. 67, no. 11, pp. 5143 – 5153, Nov. 2001.

[5] “TAMU Oceanography: Gas Hydrates,” http://www-

ocean.tamu.edu/Quarterdeck/QD5.3/sassen.html, Accessed 14 June 2004. [6] K. Kvenvolden, “Gas hydrates - Geological Perspective and Global Change”

Reviews of Geophysics, vol. 31, no. 2, pp. 173-187, May 1993. [7] D. Adam, “Fire from ice,” Nature, vol. 415, pp. 913 – 914, 29 Aug. 2002. [8] M.D. Max and W.P. Dillon, “Oceanic Methane Hydrate: The Character of the

Blake Ridge Hydrate Stability Zone, and the Potential for Methane Hydrate Extraction,” J. Petr. Geol., vol. 21, no. 3, pp. 343 – 357, July 1998.

[9] C.N. Murray, et al., “Permanent Storage of Carbon Dioxide in the Marine

Environment: The Solid CO2 Penetrator,” Energy Convers. Mgmt., vol. 37, No. 6 – 8, pp. 1067 – 1072, Aug. 1996.

[10] J.S. Gudmundsson, “Method for Production of Gas Hydrate for Transportation

and Storage,” U.S. Patent 5 536 893, 1996. [11] R.E. Rogers, “Natural Gas Storage Project,” DOE Contract DE-AC26-

97FT33203, US Department of Energy, 1997. [12] W.L. Mao, et al., “Hydrogen Clusters in Clathrate Hydrates,” Science, vol. 297,

pp. 2247 – 2249, Sept. 2002.

Page 122: Examination of the Effects of Biosurfactant Concentration

-108-

[13] G.Y. Yevi and R.E. Rogers, “Storage of Fuels in Hydrates for Natural Gas Vehicles (NGVs),” J. Energy Res. Tech., vol. 118, pp. 209 – 213, Sept. 1996.

[14] A.J. Barduhn, H.E. Towlson, and Y.C. Hu, “The Properties of Some New Gas

Hydrates and Their Use in Demineralizing Sea Water,” AICHE Journal, vol. 8, no. 2, pp. 176 – 182, May 1962.

[15] E.G. Hammerschmidt, “Gas Hydrate Formations in Natural Gas Pipe Lines,” Oil

Gas J., vol. 37, no. 52, pp. 66 – 72, May 1939. [16] Z. Huo, et al., “Hydrate plug prevention by anti-agglomeration,” Chem. Eng. Sci.,

vol. 56, no. 17, pp. 4979 – 4991, Sept. 2001. [17] T. Appenzeller, “Fire and Ice Under the Deep-Sea Floor,” Science, vol. 252, pp.

1790 – 1792, June 1991. [18] S.P. Hesselbo, et al., “Massive dissociation of gas hydrate during a Jurassic

oceanic anoxic event,” Nature, vol. 406, pp. 392 – 395, July 2000. [19] E.D. Sloan, “Fundamental principles and applications of natural gas hydrates,”

Nature, vol. 426, pp. 353 – 359, 20 Nov. 2003. [20] S.B. Jacobsen, “Gas hydrates and deglaciations,” Nature, vol. 412, pp. 691 – 693,

16 Aug. 2001. [21] M.J. Kennedy, N. Christie-Blick, and L.E. Sohl, “Are Proterozoic cap carbonates

and isotopic excursions a record of gas hydrate destabilization following Earth’s coldest intervals?” Geology, vol. 29, no. 5, pp. 443 – 446, May 2001.

[22] J.A. Ripmeester and C.I. Ratcliffe, “129Xe NMR Studies of Clathrate Hydrates:

New Guests for Structure II and Structure H,” J. Phys. Chem., vol. 94, no. 25, pp. 8773 – 8776, 13 Dec. 1990.

[23] Y.F. Makogon, Hydrates of Natural Gas Tulsa: PennWell, 1981. [24] L.M. Prescott, et al., Microbiology, 5ed. Boston: McGraw-Hill, 2002. [25] N. Kosaric, “Biosurfactants in industry,” Pure and App. Chem., vol. 64, no. 11,

pp. 1731 – 1737, Nov. 1992. [26] C.R. Kothapalli, “Catalysis of gas hydrates by biosurfactants in seawater-

saturated sand/clay,” M.S. Thesis, Mississippi State University, CHE Dept., Starkville, MS, 2002.

Page 123: Examination of the Effects of Biosurfactant Concentration

-109-

[27] A.J. Hill and S. Ghoshal, “Micellar Solubilization of Naphthalene and Phenanthrene from Nonaqueous-Phase Liquids,” Env. Sci. Tech., vol. 36, no. 18, pp. 3901 – 3907, 15 Sept. 2002.

[28] S. Lang and D. Wullbrandt, “Rhamnose lipids – biosynthesis, microbial

production and application potential,” Appl. Microbiol. Biotechnol., vol. 51, pp. 22 – 32, 1999.

[29] S. Harvey, et al., “Enhanced Removal of Exxon Valdez Spilled Oil From Alaskan

Gravel by a Microbial Surfactant,” Biotechnol., vol. 8, pp. 228 – 230, March 1990.

[30] Y. Zhong and R.E. Rogers, “Surfactant effects on gas hydrate formation,” Chem.

Eng. Sci., vol. 55, no. 19, pp. 4175 – 4187, Oct. 2000. [31] M.S. Lee, “The effects of biosurfactants on gas hydrate formation in ocean

sediments,” M.S. Thesis, Mississippi State University, CHE Dept., Starkville, MS, 2001.

[32] S. Thangamani and G.S. Shreve, “Effect of Anionic Biosurfactant on Hexadecane

Partitioning in Multiphase Systems,” Env. Sci. Tech., vol. 28, no. 12, pp. 1993 – 2000, Nov. 1994.

[33] J.T. Champion, et al., “Electron Microscopy of Rhamnolipid (Biosurfactant)

Morphology: Effects of pH, Cadmium, and Octadecane,” J. Coll. Interf. Sci., vol. 170, no. 2, pp. 569 – 574, 15 Mar. 1995.

[34] “Biodegradable Detergents,”

http://www.natick.army.mil/soldier/media/fact/ss&t/Biodegradable_Detergents.PDF, Accessed 10 May 2004.

[35] E. Rosenberg, “Microbial diversity as a source of useful biopolymers,” J. Ind.

Microbio., vol. 11, no. 3, pp. 131 – 137, May 1993. [36] P.D. Dholabhai, N. Kalogerakis, and P.R. Bishnoi, “Kinetics of Methane Hydrate

Formation in Aqueous Electrolyte Solutions,” Can. J. Chem. Eng., vol. 71, pp. 68 – 74, Feb. 1993.

[37] B. Tohidi, A. Danesh, and A. Todd, “Predicting pipeline hydrate formation,” The

Chem. Eng., iss. 642, pp. 32 – 37, 25 Sept. 1997. [38] M.D. Jager, C.J. Peters, and E.D. Sloan, “Experimental determination of methane

hydrate stability in methanol and electrolyte solutions,” Fluid Phase Eq., vol. 193, iss. 1 – 2, pp. 17 – 28, 30 Jan. 2002.

Page 124: Examination of the Effects of Biosurfactant Concentration

-110-

[39] D.D. Link, et al., “Formation and dissociation studies for optimizing the uptake of methane by methane hydrates,” Fluid Phase Eq., vol. 211, iss. 1, pp. 1 – 10, 30 Aug. 2003.

[40] G.D. Ginsburg, et al., “24. Sediment Grain-Size Control on Gas Hydrate

Presence, Sites 994, 995, and 997,” in Proc. of the Ocean Drilling Program, Scientific Results, vol. 164, pp. 237 – 245, May 2000.

[41] S.B. Cha, et al., “A Third-Surface Effect on Hydrate Formation,” J. Phys. Chem.,

vol. 92, no. 23, pp. 6492 – 6494, 17 Nov. 1988. [42] R.E. Rogers, et al., “Catalysis of Gas Hydrates by Biosurfactants in Seawater-

Saturated Sand/Clay,” Can. J. Chem. Eng., vol. 81, pp. 973 – 980, Oct. 2003. [43] T. Uchida, T. Ebinuma, and T. Ishizaki, “Dissociation Condition Measurements

of Methane Hydrate in Confined Small Pores of Porous Glass,” J. Phys. Chem. B, vol. 103, no. 18, pp. 3659 – 3662, 6 May 1999.

[44] H.O. Kono and B. Budhijanto, “Modeling of Gas Hydrate Formation by

Controlling the Interfacial Boundary Surfaces,” in Proceedings of the Fourth International Conference on Gas Hydrates, Yokohama, Japan, pp. 461 – 464, May 19 – 23, 2002.

[45] J.K. Mitchell, Fundamentals of Soil Behavior, 2ed. NY: John Wiley and Sons,

1993. [46] J.F. Banfield and K.H. Nealson, “Geomicrobiology: Interactions Between

Microbes and Minerals,” Rev. Mineralogy, vol. 35, pp. 81 – 122, 2004. [47] H.L. Bohn, et al., Soil Chemistry, NY: John Wiley and Sons, 1979. [48] D.G. Vilenskii, Soil Science, Translated by A. Birron and Z.S. Cole, 3ed.

Moscow: State Teachers’ College, 1957. [49] S.M. Lipson and G. Stotzky, “Specificity of virus adsorption to clay minerals,”

Can. J. Microbio., vol. 31, no. 1, pp. 50 – 52, Jan. 1985. [50] J. Grandjean, j. Bujdák, and P. Komadel, “NMR study of surfactant molecules

intercalated in montmorillonite and in silylated montmorillonite,” Clay Minerals, vol. 38, no. 3, pp. 367 – 373, Sept. 2003.

[51] “Sources. Negative Charge:,” http://jan.ucc.nau.edu/~doetqp-

p/courses/env320/lec13/Lec13.html, Accessed 12 May 2004.

Page 125: Examination of the Effects of Biosurfactant Concentration

-111-

[52] “Solvay Precipitate Calcium Carbonate – Physical Properties,” http://www.solvaypcc.com/safety_environment/0,5261,1000029-_EN,00.html, Accessed 12 May 2004.

[53] “243 Notes Week 10a,”

http://www.usask.ca/geology/classes/geol243/243notes/243week10a.html, Accessed 17 June 2004.

[54] D.J. Shaw, Introduction to Colloid and Surface Chemistry, 3rd ed. London:

Butterworths, 1980. [55] K.L. Mittal, Micellization, Solubilization, and Microemulsions, Vol. 1, NY:

Plenum Press, 1977. [56] “KCPC Education Resource Website: 9.5.5 – Anionic Surfactants,”

http://www.kcpc.usyd.edu.au/discovery/9.5.5-short/9.5.5_anionic.html, Accessed 14 June 2004.

[57] “-- 4: Surface Chemistry --,” http://www.erpt.org/032Q/Nelsc-01.pdf, Accessed

14 June 2004. [58] “KCPC Education Resource Website: 9.5.5 – Cationic Surfactants,”

http://www.kcpc.usyd.edu.au/discovery/9.5.5-short/9.5.5_cationic.html, Accessed 14 June 2004.

[59] “Surfactants ~ Oleochemicals – China GreatVista Chemicals,”

http://www.greatvistachemicals.com/surfactants_and_oleochemicals/, Accessed 14 June 2004.

[60] “KCPC Education Resource Website: 9.5.5 – Nonionic Surfactants,”

http://www.kcpc.usyd.edu.au/discovery/9.5.5-short/9.5.5_nonionic.html, Accessed 14 June 2004.

[61] E.Z. Ron, E. Rosenberg, “Natural roles of biosurfactants,” Envir. Microbio., vol.

3, no. 4, pp. 229 – 236, 2001. [62] R.M. Miller, “Biosurfactant-facilitated Remediation of Metal-contaminated

Soils,” Environmental Health Perspectives, vol. 103, suppl. 1, pp. 59 – 62, Feb. 1995.

[63] M.U. Jianhai, et al., “Formation of wormlike micelles in anionic surfactant AES

aqueous solutions,” Chinese Science Bulletin, vol. 46, no. 16, pp. 1360 – 1364, Aug. 2001.

Page 126: Examination of the Effects of Biosurfactant Concentration

-112-

[64] Z.A. Al-Anber, M.A. Floriano, and A.D. Mackie, “Sphere-to-rod transitions of micelles in model nonionic surfactant solutions,” J. Chem. Phys., vol. 118, no. 8, pp. 3816 – 3826, Feb. 2003.

[65] S. Roy, A. Mehra, and D. Bhowmick, “Prediction of Solubility of Nonpolar Gases

in Micellar Solutions of Ionic Surfactants,” J. Coll. Interf. Sci., vol. 196, no. 1, pp. 53 – 61, 1 Dec. 1997.

[66] R.A. Al-Tahhan, et al., “Rhamnolipid-Induced Removal of Lipopolysaccharide

from Pseudomonas aeruginosa: Effect on Cell Surface Properties and Interaction with Hydrophobic Substrates,” App. Envir. Microbio., vol. 66, no. 8, pp. 3262 – 3268, Aug. 2000.

[67] N. Li, S. Liu, and H. Luo, “A New Method for the Determination of the First and

Second CMC in CTAB Solution by Resonance Rayleigh Scattering Technology,” Analytical Letters, vol. 35, no. 7, pp. 1229 – 1238, May 2002.

[68] D.L. Gutnick, “The Emulsan Polymer: Perspectives on a Microbial Capsule as an

Industrial Product,” Biopolymers, 26, pp. S223 – S240, 1987. [69] M.E. Sumner, Handbook of Soil Science, Boca Raton, FL: CRC Press, 2000. [70] “ENSC 4734 Topic VI Sorption Phenomena,”

http://soils1.cses.vt.edu/MJE/ENSC4734/Slides/Topic6.pdf, Accessed 29 May 2004.

[71] A. Vysniauskas and P.R. Bishnoi, “A Kinetic Study of Methane Hydrate

Formation,” Chem. Eng. Sci., vol. 38, no. 7, pp. 1061 – 1072, July 1983. [72] V.M. Starov, S.A. Zhdanov, and M.G. Velarde, “Capillary imbibition of

surfactant solutions in porous media and thin capillaries: partial wetting case,” J. Coll. and Interf. Sci., vol. 273, no.2, pp. 589 – 595, 15 May 2004.

[73] D. Kashchiev and A. Firoozabadi, “Induction time in crystallization of gas

hydrates,” J. Crystal Growth, vol. 250, iss. 3 – 4, pp. 499 – 515, Apr. 2003. [74] A. Vysniauskas and P.R. Bishnoi, “Kinetics Study of Ethane Hydrate Formation,”

Chem. Eng. Sci., vol. 40, no. 2, pp. 299 – 303, Feb. 1985. [75] F. Varaminian, “The Role of Heat Transfer in Kinetics of Hydrate Formation,” in

Proceedings of the Fourth International Conference on Gas Hydrates, Yokohama, Japan, pp. 465 – 468, May 19 – 23, 2002.

Page 127: Examination of the Effects of Biosurfactant Concentration

-113-

[76] M. Mork, J.S. Gudmundsson, and M. Parlaktuna, “Hydrate Formation Rate in a Continuous Stirred Tank Reactor,” presented at 2001 International Gas Research Conference, Amsterdam, The Netherlands, November 5 – 8, 2001.

[77] P. Englezos and P.R. Bishnoi, “Prediction of Gas Hydrate Formation Conditions

in Aqueous Electrolyte Solutions,” AIChE Journal, vol. 34, no. 10, pp. 1718 – 1721, Oct. 1988.

[78] J.L. Roo, et al., “Occurrence of Methane Hydrate in Saturated and Unsaturated

Solutions of Sodium Chloride and Water in Dependence of Temperature and Pressure,” AIChE Journal, vol. 29, no. 4, pp. 651 – 657, July 1983.

Page 128: Examination of the Effects of Biosurfactant Concentration

-114-

APPENDIX A

Experimental Data

Page 129: Examination of the Effects of Biosurfactant Concentration

Table A.1. Experimental Plan

Conc. Run # Surface Biosurfactant

Surf. Tension (mN/m) (ppm)

Ionic Content

Amt of Sand (g)

Amt of Bentonite (g)

Amt of Kaolinite (g)

Amt of Nontronite (g)

Amt of Carbonate (g)

100 Sand Rhamnolipid 27.0 1000 DI Water 70 0 0 0 0 101 Sand Rhamnolipid 29.8 100 DI Water 70 0 0 0 0 102 Sand Rhamnolipid 54.6 10 DI Water 70 0 0 0 0 103 Sand None 73.8 0 DI Water 70 0 0 0 0 104 Sand/Bent Rhamnolipid 26.8 1000 DI Water 67 3 0 0 0 105 Sand/Bent Rhamnolipid 30.2 100 DI Water 67 3 0 0 0 106 Sand/Bent Rhamnolipid 54.6 10 DI Water 67 3 0 0 0 107 Sand/Bent None 72.8 0 DI Water 67 3 0 0 0 108 Sand/Kaol Rhamnolipid 27.2 1000 DI Water 67 0 3 0 0 109 Sand/Kaol Rhamnolipid 29.9 100 DI Water 67 0 3 0 0 110 Sand/Kaol Rhamnolipid 46.2 10 DI Water 67 0 3 0 0 111 Sand/Kaol None 72.9 0 DI Water 67 0 3 0 0 112 Sand Emulsan 40.7 1000 DI Water 70 0 0 0 0 113 Sand Emulsan 44.4 100 DI Water 70 0 0 0 0 114 Sand Emulsan 62.2 10 DI Water 70 0 0 0 0 115 Sand/Bent Emulsan 40.6 1000 DI Water 67 3 0 0 0 116 Sand/Bent Emulsan 43.9 100 DI Water 67 3 0 0 0 117 Sand/Bent Emulsan 56.5 10 DI Water 67 3 0 0 0 118 Sand/Kaol Emulsan 41.9 1000 DI Water 67 0 3 0 0 119 Sand/Kaol Emulsan 44.9 100 DI Water 67 0 3 0 0 120 Sand/Kaol Emulsan 62.1 10 DI Water 67 0 3 0 0 121 Sand None 73.3 0 Seawater 70 0 0 0 0 122 Sand Rhamnolipid 26.3 1000 Seawater 70 0 0 0 0 125 Sand/Nont None 73.3 0 DI Water 67 0 0 3 0 126 Sand/Carb None 73.4 0 DI Water 67 0 0 0 3 127 Sand/Nont Rhamnolipid 27.3 1000 DI Water 67 0 0 3 0 128 Sand/Carb Rhamnolipid 26.9 1000 DI Water 67 0 0 0 3

-115-

Page 130: Examination of the Effects of Biosurfactant Concentration

Table A.1. Experimental Plan (Cont’d)

Conc. Run # Surface Biosurfactant

Surf. Tension (mN/m) (ppm)

Ionic Content

Amt of Sand (g)

Amt of Bentonite (g)

Amt of Kaolinite (g)

Amt of Nontronite (g)

Amt of Carbonate (g)

130 Sand Rhamnolipid 27.5 1000 DI Water 70 0 0 0 0 131 Sand Rhamnolipid 30.0 100 DI Water 70 0 0 0 0 132 Sand Rhamnolipid 54.4 10 DI Water 70 0 0 0 0 133 Sand None 72.3 0 DI Water 70 0 0 0 0 134 Sand/Bent Rhamnolipid 27.1 1000 DI Water 67 3 0 0 0 135 Sand/Bent Rhamnolipid 30.3 100 DI Water 67 3 0 0 0 136 Sand/Bent Rhamnolipid 49.9 10 DI Water 67 3 0 0 0 137 Sand/Bent None 73.4 0 DI Water 67 3 0 0 0 138 Sand/Kaol Rhamnolipid 27.4 1000 DI Water 67 0 3 0 0 139 Sand/Kaol Rhamnolipid 30.2 100 DI Water 67 0 3 0 0 140 Sand/Kaol Rhamnolipid 58.5 10 DI Water 67 0 3 0 0 141 Sand/Kaol None 73.2 0 DI Water 67 0 3 0 0 142 Sand Emulsan 39.7 1000 DI Water 70 0 0 0 0 143 Sand Emulsan 44.7 100 DI Water 70 0 0 0 0 144 Sand Emulsan 62.8 10 DI Water 70 0 0 0 0 145 Sand/Bent Emulsan 39.2 1000 DI Water 67 3 0 0 0 146 Sand/Bent Emulsan 45.1 100 DI Water 67 3 0 0 0 147 Sand/Bent Emulsan 56.8 10 DI Water 67 3 0 0 0 148 Sand/Kaol Emulsan 38.6 1000 DI Water 67 0 3 0 0 149 Sand/Kaol Emulsan 44.8 100 DI Water 67 0 3 0 0 150 Sand/Kaol Emulsan 63.4 10 DI Water 67 0 3 0 0 151 Sand None 73.0 0 Seawater 70 0 0 0 0 152 Sand Rhamnolipid 26.8 1000 Seawater 70 0 0 0 0 155 Sand/Nont None 72.8 0 DI Water 67 0 0 3 0 156 Sand/Carb None 72.6 0 DI Water 67 0 0 0 3 157 Sand/Nont Rhamnolipid 27.4 1000 DI Water 67 0 0 3 0

-116-

Page 131: Examination of the Effects of Biosurfactant Concentration

Table A.1. Experimental Plan (Cont’d)

Conc. Run # Surface Biosurfactant

Surf. Tension (mN/m) (ppm)

Ionic Content

Amt of Sand (g)

Amt of Bentonite (g)

Amt of Kaolinite (g)

Amt of Nontronite (g)

Amt of Carbonate (g)

158 Sand/Carb Rhamnolipid 27.3 1000 DI Water 67 0 0 0 3 160 Sand Rhamnolipid 27.0 500 DI Water 70 0 0 0 0 161 Sand Rhamnolipid 27.5 500 DI Water 70 0 0 0 0 162 Sand Emulsan 41.2 500 DI Water 70 0 0 0 0 163 Sand Emulsan 42.1 500 DI Water 70 0 0 0 0 164 Sand/Bent Rhamnolipid 27.6 500 DI Water 67 3 0 0 0 165 Sand/Bent Rhamnolipid 27.3 500 DI Water 67 3 0 0 0 166 Sand/Bent Emulsan 41.1 500 DI Water 67 3 0 0 0 167 Sand/Bent Emulsan 41.8 500 DI Water 67 3 0 0 0 168 Sand/Kaol Rhamnolipid 27.1 500 DI Water 67 0 3 0 0 169 Sand/Kaol Rhamnolipid 27.5 500 DI Water 67 0 3 0 0 172 Sand Emulsan 45.8 100 DI Water 70 0 0 0 0 173 Sand/Kaol Rhamnolipid 27.0 1000 DI Water 67 0 3 0 0 174 Sand None 72.9 0 DI Water 70 0 0 0 0 175 Sand Rhamnolipid 49.2 10 DI Water 70 0 0 0 0 176 Sand Emulsan 43.8 100 DI Water 70 0 0 0 0 177 Sand Emulsan 61.9 10 DI Water 70 0 0 0 0 179 Sand Rhamnolipid 27.1 1000 DI Water 70 0 0 0 0 180 Sand/Bent Rhamnolipid 29.7 100 DI Water 67 3 0 0 0 181 Sand/Kaol Rhamnolipid 27.5 1000 DI Water 67 0 3 0 0 182 Sand/Kaol Rhamnolipid 26.5 500 DI Water 67 0 3 0 0 183 Sand/Bent Emulsan 39.2 1000 DI Water 67 3 0 0 0 184 Sand Rhamnolipid 26.1 1000 Seawater 70 0 0 0 0 185 Sand Emulsan 40.1 500 Seawater 70 0 0 0 0

-117-

Page 132: Examination of the Effects of Biosurfactant Concentration

-118-

Table A.2. Surface Tension of Rhamnolipid at Room & Hydrate Temperature

Rhamnolipid/Distilled Water

Concentration (ppm)

Surf. Tension @ 35oF (mN/m)

Surf. Tension @ Room Temp.(mN/m)

0 65.7 70.2 10 49.9 46.9 20 46.3 42.0 30 43.9 39.9 40 41.3 37.3 50 41.2 35.4 60 38.2 35.3 70 36.1 33.2 80 36.4 32.1 90 35.7 31.4 100 34.0 31.7 250 31.0 29.5 500 30.4 28.8 1000 29.4 28.3

Table A.3. Surface Tension of Emulsan at Room Temperature

Emulsan/Distilled Water

Concentration (ppm)

Surf. Tension @ Room Temp. (mN/m)

0 73.4 10 55.1 20 53.7 30 47.8 40 47.4 50 47.7 60 43.3 70 43.2 80 43.1 90 43.4 100 43.7 200 41.1 300 41.2 400 41.2 500 40.8 1000 40.4

Page 133: Examination of the Effects of Biosurfactant Concentration

-119-

25.0

30.0

35.0

40.0

45.0

50.0

55.0

60.0

65.0

70.0

75.0

0 200 400 600 800 1000

Concentration Rhamnolipid (ppm)

Surf

ace

Tens

ion

(mN

/m)

Rhamn./DW @RT Rhamn./DW @35oF

Figure A.1. CMC of Rhamnolipid at Room & Refrigerated Temperature

35.0

40.0

45.0

50.0

55.0

60.0

65.0

70.0

75.0

0 200 400 600 800 1000

Concentration Rhamnolipid (ppm)

Surf

ace

Tens

ion

(mN

/m)

Emulsan @RT

Figure A.2. ST Vs Concentration of Emulsan at Room Temperature

Page 134: Examination of the Effects of Biosurfactant Concentration

Table A.4. Heat & Mass Transfer Effects on Hydrate Formation (Fig. 5.10) Run # Surface Surfactant Surf. Tension Conc. pH Induction Time Peak Formation Rate

(mN/m) (ppm) (hrs.) (mmol/min) 25 Sand/Bent None 70.3 0 -- 24.10 40.03 32 Sand/Bent None 73.3 0 -- 0.95 29.35 37 Sand/Bent None 73.4 0 22.58 66.35 13 Sand/Bent Rhamno 56.4 10 6.90 6.53 34.16 14 Sand/Bent Rhamno 57.7 10 6.70 1.30 26.16 15 Sand/Bent Rhamno 54.9 10 -- 1.50 28.17 4 Sand/Bent Rhamno 33.2 100 -- 0.60 12.22 5 Sand/Bent Rhamno 32.8 100 -- 0.77 12.14 6 Sand/Bent Rhamno 33.1 100 -- 0.70 18.02 8 Sand/Bent Rhamno 28.3 1000 6.49 0.95 7.61

11 Sand/Bent Rhamno 28.3 1000 6.54 5.58 10.65 12 Sand/Bent Rhamno 28.0 1000 6.46 7.43 8.35

Table A.5. Effect of Rhamnolipid on Ottawa Sand, Averaged (Fig. 5.6)

Rhamnolipid

Concentration (ppm)Ottawa Sand, Peak Form.

Rate (mmol/min) STDEV 0 0.54 0.04

10 0.49 #DIV/0! 100 0.80 0.20 500 1.62 0.00

1000 1.64 0.32

-120-

Page 135: Examination of the Effects of Biosurfactant Concentration

-121-

Table A.6. Effect of Rhamnolipid on Ottawa Sand/Bentonite, Averaged (Fig. 5.7)

Rhamnolipid

Concentration (ppm) OS/Bent. Peak Form.

Rate (mmol/min) STDEV

0 0.90 0.25 10 0.93 0.25

100 1.89 0.35 500 2.31 0.05

1000 1.50 0.00 Table A.7. Effect of Rhamnolipid on Ottawa Sand/Kaolinite, Averaged (Fig. 5.8)

Rhamnolipid

Concentration (ppm) OS/Kaolin Peak Form.

Rate (mmol/min) STDEV

0 0.45 0.04 10 0.48 0.06

100 0.50 0.02 500 1.16 0.42

1000 1.70 0.33 Table A.8. Effect of Emulsan on Ottawa Sand, Averaged (Fig. 5.4)

Emulsan

Concentration (ppm) Ottawa Sand, Peak Form.

Rate (mmol/min) STDEV

0 0.54 0.04 10 0.83 0.07

100 0.62 #DIV/0! 1000 1.49 0.11

Page 136: Examination of the Effects of Biosurfactant Concentration

-122-

Table A.9. Effect of Emulsan on Ottawa Sand/Bentonite, Averaged (Fig. 5.4)

Emulsan

Concentration (ppm) OS/Bent. Peak Form. Rate

(mmol/min) STDEV

0 0.90 0.25 10 0.98 0.04

100 1.07 0.12 500 1.80 0.25

1000 1.93 0.66 Table A.10. Effect of Emulsan on Ottawa Sand/Kaolinite, Averaged (Fig. 5.4)

Emulsan

Concentration (ppm) OS/Kaolin Peak Form. Rate

(mmol/min) STDEV

0 0.45 0.04 10 0.36 0.00

100 0.43 #DIV/0! 1000 0.39 0.04

Page 137: Examination of the Effects of Biosurfactant Concentration

Table A.11. Effect of Rhamnolipid on Varied Surfaces, Averaged (Fig. 5.3, 5.5)

Ottawa Sand OS/Bentonite OS/Kaolinite OS/Nontronite OS/Aragonite

Concentration (ppm)

Peak Form. Rate

(mmol/min) STDEV

Peak Form. Rate

(mmol/min) STDEV

Peak Form. Rate

(mmol/min) STDEV

Peak Form. Rate

(mmol/min) STDEV

Peak Form. Rate

(mmol/min) STDEV 0 0.54 0.04 0.90 0.25 0.45 0.04 0.40 0.13 0.75 0.00

1000 1.64 0.32 1.50 0.00 1.70 0.33 1.59 0.37 1.42 0.15 Table A.12. Compiled Experimental Data

Surf. Tension Conc. Ind. TimeInitial Hydrate

Form. Rate Average-to-Peak

Form. Rate Peak Form.

Rate ∆ntot Run # Surface Surfactant (mN/m) (ppm) (hrs.) (mmol/min) (mmol/min) (mmol/min) (mmoles) 100 Sand Rhamnolipid 27.0 1000 2.53 0.86 1.16 1.99 81.63 130 Sand Rhamnolipid 27.5 1000 13.18 0.90 0.80 1.35 88.41 179 Sand Rhamnolipid 27.1 1000 11.40 0.74 0.93 1.59 90.03 160 Sand Rhamnolipid 27.0 500 75.38 1.04 0.66 1.62 90.23 161 Sand Rhamnolipid 27.5 500 5.89 0.99 0.96 1.62 90.83 101 Sand Rhamnolipid 29.8 100 1.79 0.44 0.30 0.66 31.16 131 Sand Rhamnolipid 30.0 100 1.20 0.09 0.33 0.94 91.09 102 Sand Rhamnolipid 54.6 10 >48 n/a n/a n/a 5.48 132 Sand Rhamnolipid 54.4 10 6.67 0.08 0.11 0.49 56.68 175 Sand Rhamnolipid 49.2 10 >48 n/a n/a n/a n/a 103 Sand None 73.8 0 >48 n/a n/a n/a 0.00 133 Sand None 72.3 0 11.63 0.23 0.17 0.56 78.25

-123-

Page 138: Examination of the Effects of Biosurfactant Concentration

Table A.12. Compiled Experimental Data (Cont’d)

Surf. Tension Conc. Ind. TimeInitial Hydrate

Form. Rate Average-to-Peak

Form. Rate Peak Form.

Rate ∆ntot Run # Surface Surfactant (mN/m) (ppm) (hrs.) (mmol/min) (mmol/min) (mmol/min) (mmoles)174 Sand None 72.9 0 3.90 0.06 0.23 0.51 77.58 104 Sand/Bent Rhamnolipid 26.8 1000 3.74 0.71 0.71 1.50 88.30 134 Sand/Bent Rhamnolipid 27.1 1000 25.79 0.79 0.72 1.50 91.06 164 Sand/Bent Rhamnolipid 27.6 500 1.15 1.26 1.08 2.27 91.93 165 Sand/Bent Rhamnolipid 27.3 500 2.11 1.99 0.99 2.34 89.56 105 Sand/Bent Rhamnolipid 30.2 100 1.62 1.37 1.20 2.27 92.06 135 Sand/Bent Rhamnolipid 30.3 100 1.33 1.16 0.92 1.79 89.19 180 Sand/Bent Rhamnolipid 29.7 100 3.98 n/a n/a 1.60 93.37 106 Sand/Bent Rhamnolipid 54.6 10 4.88 0.55 0.53 0.75 58.88 136 Sand/Bent Rhamnolipid 49.9 10 7.98 n/a n/a 1.11 77.74 107 Sand/Bent None 72.8 0 0.96 0.50 0.44 0.72 57.53 137 Sand/Bent None 73.4 0 3.68 0.70 0.68 1.08 77.88 108 Sand/Kaol Rhamnolipid 27.2 1000 2.52 1.19 0.76 1.89 81.45 138 Sand/Kaol Rhamnolipid 27.4 1000 >48 n/a n/a n/a 0.00 173 Sand/Kaol Rhamnolipid 27.0 1000 1.80 0.89 0.81 1.32 80.44 181 Sand/Kaol Rhamnolipid 27.5 1000 1.50 0.82 0.89 1.88 74.09 168 Sand/Kaol Rhamnolipid 27.1 500 7.93 0.26 0.32 0.67 78.96 169 Sand/Kaol Rhamnolipid 27.5 500 3.53 0.80 0.77 1.36 79.54 182 Sand/Kaol Rhamnolipid 26.5 500 3.37 0.27 0.51 1.44 86.29 109 Sand/Kaol Rhamnolipid 29.9 100 1.79 0.09 0.13 0.51 72.91 139 Sand/Kaol Rhamnolipid 30.2 100 4.67 0.14 0.02 0.48 91.05 110 Sand/Kaol Rhamnolipid 46.2 10 5.46 0.04 0.09 0.52 53.87 140 Sand/Kaol Rhamnolipid 58.5 10 2.35 n/a n/a 0.43 67.90 111 Sand/Kaol None 72.9 0 2.18 0.02 0.03 0.42 50.89 141 Sand/Kaol None 73.2 0 5.28 n/a n/a 0.48 70.19

-124-

Page 139: Examination of the Effects of Biosurfactant Concentration

Table A.12. Compiled Experimental Data (Cont’d)

Surf. Tension Conc. Ind. TimeInitial Hydrate

Form. Rate Average-to-Peak

Form. Rate Peak Form.

Rate ∆ntot Run # Surface Surfactant (mN/m) (ppm) (hrs.) (mmol/min) (mmol/min) (mmol/min) (moles) 112 Sand Emulsan 40.7 1000 6.58 0.23 0.73 1.41 85.02 142 Sand Emulsan 39.7 1000 23.10 0.28 0.76 1.56 89.79 162 Sand Emulsan 41.2 500 6.63 0.18 0.13 0.43 10.21 163 Sand Emulsan 42.1 500 14.68 0.11 0.07 0.43 6.61 185 Sand Emulsan 40.1 500 11.09 n/a n/a 0.43 16.27 113 Sand Emulsan 44.4 100 >48 n/a n/a n/a 15.63 143 Sand Emulsan 44.7 100 >48 n/a n/a n/a 0.00 172 Sand Emulsan 45.8 100 1.51 0.29 0.28 0.62 54.32 176 Sand Emulsan 43.8 100 >48 n/a n/a n/a 0.00 114 Sand Emulsan 62.2 10 >48 n/a n/a n/a 3.73 144 Sand Emulsan 62.8 10 0.99 0.11 0.24 0.78 67.00 177 Sand Emulsan 61.9 10 12.18 n/a n/a 0.88 64.18 115 Sand/Bent Emulsan 40.6 1000 4.47 0.74 1.19 2.68 87.09 145 Sand/Bent Emulsan 39.2 1000 1.58 0.35 0.81 1.47 90.49 183 Sand/Bent Emulsan 39.2 1000 9.47 n/a n/a 1.63 85.26 166 Sand/Bent Emulsan 41.1 500 3.78 1.06 1.22 1.98 88.18 167 Sand/Bent Emulsan 41.8 500 2.23 0.73 1.12 1.62 85.78 116 Sand/Bent Emulsan 43.9 100 3.33 0.64 0.58 0.98 70.43 146 Sand/Bent Emulsan 45.1 100 1.13 0.82 0.69 1.15 82.29 117 Sand/Bent Emulsan 56.5 10 5.74 n/a n/a 1.01 81.31

-125-

Page 140: Examination of the Effects of Biosurfactant Concentration

Table A.12. Compiled Experimental Data (Cont’d)

Surf.

Tension Conc. Ind. TimeInitial Hydrate

Form. Rate Average-to-Peak

Form. Rate Peak Form.

Rate ∆ntot Run # Surface Surfactant (mN/m) (ppm) (hrs.) (mmol/min) (mmol/min) (mmol/min) (moles) 147 Sand/Bent Emulsan 56.8 10 1.47 n/a n/a 0.95 85.86 118 Sand/Kaol Emulsan 41.9 1000 3.98 0.03 0.05 0.36 53.21 148 Sand/Kaol Emulsan 38.6 1000 2.98 n/a n/a 0.42 68.17 119 Sand/Kaol Emulsan 44.9 100 10.88 n/a n/a 0.43 19.12 149 Sand/Kaol Emulsan 44.8 100 >48 n/a n/a n/a 0.00 120 Sand/Kaol Emulsan 62.1 10 32.23 n/a n/a 0.36 16.07 150 Sand/Kaol Emulsan 63.4 10 >48 n/a n/a 0.36 17.94 121 Sand None/SW 73.3 0 4.73 0.05 0.04 0.42 55.11 151 Sand None/SW 73.0 0 6.52 0.03 0.03 0.36 62.73 122 Sand Rhamnolipid/SW 26.3 1000 3.39 0.91 0.63 1.24 73.37 152 Sand Rhamnolipid/SW 26.8 1000 5.34 0.74 0.30 1.93 74.78 184 Sand Rhamnolipid/SW 26.1 1000 3.03 0.62 0.61 1.24 71.05 125 Sand/Nont None 73.3 0 9.19 0.03 0.07 0.30 33.05 155 Sand/Nont None 72.8 0 >48 n/a n/a 0.49 12.34 126 Sand/Carb None 73.4 0 0.43 0.42 0.31 0.75 74.12 156 Sand/Carb None 72.6 0 0.77 0.37 0.39 0.75 73.29 127 Sand/Nont Rhamnolipid 27.3 1000 10.13 1.32 1.27 1.85 85.62 157 Sand/Nont Rhamnolipid 27.4 1000 2.62 1.15 0.69 1.33 84.90 128 Sand/Carb Rhamnolipid 26.9 1000 0.70 0.63 0.78 1.31 88.95 158 Sand/Carb Rhamnolipid 27.3 1000 33.93 0.58 0.81 1.52 85.49

-126-

Page 141: Examination of the Effects of Biosurfactant Concentration

-127-

APPENDIX B

Peng-Robinson Calculations

Page 142: Examination of the Effects of Biosurfactant Concentration

-128-

Given: Initial Temperature, Ti = 274.83K Final Temperature, Tf = 274.90K Initial Pressure, Pi = 19.95 atm Final Pressure, Pf = 19.90 atm Volume, V = 440 ml Time, ∆t = 30 sec. Critical Temperature, Tc = 204.656 K Critical Pressure, Pc = 45.43 atm Acentric Factor, ω = 0.01916

( ) ( ) ( )32223 2310)( BBABzBBAzBzzf −−−⋅−−+⋅+−==

2

)(45724.0

r

r

TPA ⋅⋅

=ωα

r

r

TPB ⋅

=0778.0

( ) ( )[ ]25.02 126992.054226.137464.01)( rTa −⋅+++= ωωω

cr P

PP =

cr T

TT =

92878.0=fz

92853.0=iz

⋅−

⋅⋅

=

−⋅=∆

274.83K0.9285395.19

274.90K0.9287890.19

08206.0

440.0 atmatm

KmolatmL

LTz

PTz

PRVn

ii

i

ff

f

mmolmoleEn 34.1334.1 =−=∆

min/68.2min5.0

34.1 mmolmmoltnrformation ==

∆∆

=−