a field application of nanoparticles for improved downhole

88
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2015-05-25 A Field Application of Nanoparticles For Improved Downhole Losses in Invert Emulsion Drilling Fluids Borisov, Alexey Borisov, A. (2015). A Field Application of Nanoparticles For Improved Downhole Losses in Invert Emulsion Drilling Fluids (Unpublished master's thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/24735 http://hdl.handle.net/11023/2268 master thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

Upload: others

Post on 05-May-2022

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A Field Application of Nanoparticles For Improved Downhole

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2015-05-25

A Field Application of Nanoparticles For Improved

Downhole Losses in Invert Emulsion Drilling Fluids

Borisov, Alexey

Borisov, A. (2015). A Field Application of Nanoparticles For Improved Downhole Losses in Invert

Emulsion Drilling Fluids (Unpublished master's thesis). University of Calgary, Calgary, AB.

doi:10.11575/PRISM/24735

http://hdl.handle.net/11023/2268

master thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

Page 2: A Field Application of Nanoparticles For Improved Downhole

UNIVERSITY OF CALGARY

A Field Application of Nanoparticles for Improved Downhole Losses in Invert Emulsion Drilling

Fluids

by

Alexey S. Borisov

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN CHEMICAL AND PETROLEUM ENGINEERING

CALGARY, ALBERTA

MAY, 2015

© Alexey S. Borisov 2015

Page 3: A Field Application of Nanoparticles For Improved Downhole

ii

Abstract

Invasion of drilling fluids filtrate and solids into porous, permeable, fractured or vuggy zones can

cause formation damage and presents a major source of drilling problems. Furthermore, downhole

mud losses also increase environmental and financial risks associated with drilling operations,

costing over $1B annually.

This thesis investigates the use of in situ prepared calcium carbonate nanoparticles (CNP)

for fluid loss prevention in invert emulsion drilling fluids. CNP at 5 wt% concentration were

synthesized within a custom ‘carrier’ emulsion using a modified microemulsion approach.

Subsequently, the carrier emulsion was used to deliver target concentration of NPs to a host drilling

fluid of interest via volumetric dilution. High pressure, high temperature (HPHT) fluid loss

experiments on commercial invert emulsion drilling fluids showed that CNP at concentration of

0.5 wt% provided a 20–50% improvement over conventional lost circulation materials (LCM). In

addition, basic properties of mud samples were not affected significantly in the presence of the

carrier emulsion.

In order to evaluate performance of CNP under real-life conditions, six full-scale field tests

were conducted in horizontal wells in Alberta, Canada. Industry-scale synthesis of CNP followed

the lab-bench process and was implemented at a specialized mixing facility. The results suggested

that the scale-up from 3×10-4 m3 (300 mL) to 20 m3 did not affect average particle size or final

properties of the carrier emulsion. Furthermore, field HPHT data showed good agreement with the

lab experiments, where the average fluid loss in the test wells was reduced by 20–30% compared

to the control wells using conventional drilling fluids. Finally, analysis of mud losses revealed that

the cumulative losses while drilling were on average 20–30% lower in the presence of 0.5 wt%

CNP, which suggested that NPs help to reduce downhole losses.

Page 4: A Field Application of Nanoparticles For Improved Downhole

iii

Acknowledgements

I would to express my deepest gratitude to my supervisor, Dr. Maen Husein, for giving me this

opportunity and for his invaluable guidance, mentorship, and patience over the years. I also would

like to thank my co-supervisor, Dr. Geir Hareland, for his helpful instruction and financial support.

I am very grateful to Mr. Jeremy Krol, Mr. David Edmonds and nFluids Inc. (Calgary, Alberta)

for all their assistance and kindness, and for making this project possible.

I would like to acknowledge the Department of Chemical and Petroleum Engineering,

NSERC, Talisman Energy, Pason Systems Corp, and Herbert and Ursula Zandmer Scholarship for

their continued financial support during my program.

For the extensive assistance during the field testing, I am especially thankful to Mr. Robert

Merkley and Mr. Trevor Jacobs (Blackstone Drilling Fluids, Calgary, Alberta) and Mr. Lorne

Simpson (Yangarra Resources, Calgary, Alberta).

Last but not least, I would like to thank Dr. Mohammad Zakaria, Dr. Oscar Contreras, Mrs.

Patricia Teichrob, Drilling Research Team and fellow graduate students for all their support and

advice.

Page 5: A Field Application of Nanoparticles For Improved Downhole

iv

Dedication

Dedicated to my loving family and all my friends, without whose motivation and inspiration this

thesis would not be possible.

Page 6: A Field Application of Nanoparticles For Improved Downhole

v

Table of Contents

Abstract ............................................................................................................................... ii

Acknowledgements ............................................................................................................ iii

Dedication .......................................................................................................................... iv

Table of Contents .................................................................................................................v

List of Tables .................................................................................................................... vii

List of Figures and Illustrations ....................................................................................... viii

List of Symbols, Abbreviations and Nomenclature .............................................................x

CHAPTER ONE: INTRODUCTION ..................................................................................1

1.1 Functions and properties of drilling fluids .................................................................1

1.2 Types of drilling fluids ..............................................................................................4

1.3 Mud circulation system ..............................................................................................5

1.4 Emulsions and surfactants .........................................................................................7

1.5 Invert emulsion drilling fluids ...................................................................................8

1.6 Mechanism of fluid loss and cross-flow filtration ...................................................11

1.7 Nanoparticles in oil and gas .....................................................................................16

1.8 Emulsion-based synthesis of nanoparticles .............................................................19

1.9 Research objectives ..................................................................................................22

CHAPTER TWO: MATERIALS AND METHODS ........................................................24

2.1 Chemicals and precursors ........................................................................................24

2.2 Drilling fluids sample preparation ...........................................................................24

2.3 Drilling fluids testing ...............................................................................................25

2.3.1 Mud weight ......................................................................................................25

2.3.2 Electrical stability ............................................................................................26

2.3.3 Rheology and gel strength ...............................................................................27

2.3.4 HPHT fluid loss test ........................................................................................28

2.3.5 Retort analysis .................................................................................................29

2.4 In situ synthesis of NPs ............................................................................................30

2.5 Carrier emulsion synthesis .......................................................................................31

2.5.1 Bench-scale synthesis of a carrier emulsion ....................................................32

2.5.2 Field-scale synthesis of a carrier emulsion ......................................................33

2.6 Particle characterization ...........................................................................................34

2.7 Field testing ..............................................................................................................35

2.7.1 Field test implementation ................................................................................37

2.7.2 Field data collection ........................................................................................38

CHAPTER THREE: RESULTS AND DISCUSSION ......................................................42

3.1 Carrier emulsion ......................................................................................................42

3.1.1 Properties of a carrier emulsion .......................................................................42

3.1.2 DLS analysis of a carrier emulsion .................................................................44

3.1.3 SEM and EDX analysis of a carrier emulsion .................................................46

3.2 Basic properties of commercial drilling fluids with CNP ........................................50

3.3 Fluid loss in commercial drilling fluids with CNP ..................................................55

3.4 Field testing ..............................................................................................................58

Page 7: A Field Application of Nanoparticles For Improved Downhole

vi

3.4.1 HPHT fluid loss ...............................................................................................59

3.4.2 The effect of concentration of carrier ..............................................................62

3.4.3 Cumulative mud losses ....................................................................................63

3.4.4 Mud losses per 100 m drilled ..........................................................................65

3.4.5 Correlation of HPHT fluid loss and mud losses ..............................................67

CHAPTER FOUR: CONCLUSIONS, CONTRIBUTIONS AND RECOMMENDATIONS

...................................................................................................................................68

4.1 Conclusions ..............................................................................................................68

4.2 Contributions to knowledge .....................................................................................71

4.3 Recommendations for future work ..........................................................................71

REFERENCES ..................................................................................................................73

Page 8: A Field Application of Nanoparticles For Improved Downhole

vii

List of Tables

Table 2-1: A summary of commercial drilling fluid lab samples. ................................................ 25

Table 2-2: A typical formulation of a carrier emulsion with 5 wt% of CNP. ............................... 32

Table 2-3: An overview of control and test wells. ........................................................................ 36

Table 3-1: Comparison of the properties of the carrier emulsion samples with 5 wt% CNP

prepared on a lab and field scale. .......................................................................................... 43

Table 3-2: Comparison of DLS data on carrier emulsion samples with 5 wt% CNP prepared

in a laboratory vs field scale. 25 °C, THF dispersant. .......................................................... 45

Page 9: A Field Application of Nanoparticles For Improved Downhole

viii

List of Figures and Illustrations

Figure 1: A schematic of a rig mud circulation system. ................................................................. 6

Figure 2: Drilling fluid filtration in the presence of conventional LCM (A) and NPs (B). .......... 15

Figure 3: A schematic of chemical coprecipitation using bulk aqueous phase (top) and a

reverse micelle approach (bottom). ....................................................................................... 21

Figure 4: A FANN Model 140 mud balance. ............................................................................... 26

Figure 5: An OFITE electrical stability tester (Model #131-50). ................................................. 26

Figure 6: A Fann Model 35 viscometer. ....................................................................................... 27

Figure 7: An OFITE HPHT filter press (Model #170-00). ........................................................... 28

Figure 8: An OFITE 50 mL retort kit (Model #165-14). .............................................................. 30

Figure 9: Surface locations of the control and test well groups A, B, C, and D. .......................... 35

Figure 10: A schematic of NPs field implementation using a carrier approach. .......................... 37

Figure 11: A diagram explaining calculation of mud losses in the field. ..................................... 39

Figure 12: Example of the discrepancy between manual and digital floating sensors in a mud

storage tank. .......................................................................................................................... 40

Figure 13: Long-term stability of carrier emulsion samples with 5 wt% CNP prepared in a

laboratory vs field scale. A typical commercial control invert is shown for comparison. ... 44

Figure 14: Intensity size distribution in a carrier emulsion samples with 5 wt% CNP

produced in a laboratory (A) vs field (B). ............................................................................. 46

Figure 15: SEM image of the profile of the filter cake formed by a lab sample of carrier

emulsion with 5 wt% CNP (A); the corresponding overlapped EDX images showing

uniform distribution of calcium, carbon, oxygen, potassium, and chloride (B). .................. 48

Figure 16: SEM images of two different areas within the profile of a filter cake formed by a

lab sample of carrier emulsion with 5 wt% CNP (A, D); the corresponding overlapped

EDX images showing correlation between calcium, carbon, and oxygen (B, E); the same

areas showing correlation between potassium, and chlorine (C, F)...................................... 49

Figure 17: Impact of 0.5 wt% CNP on mud weight of commercial invert emulsion drilling

fluids using in situ and carrier emulsion methods. ............................................................... 51

Figure 18: Impact of 0.5 wt% CNP on electrical stability of commercial invert emulsion

drilling fluids using in situ and carrier emulsion methods. ................................................... 52

Page 10: A Field Application of Nanoparticles For Improved Downhole

ix

Figure 19: Impact of 0.5 wt% CNP on plastic viscosity (A, B) and yield point (C, D) of

commercial invert emulsion drilling fluids. .......................................................................... 53

Figure 20: Impact of 0.5 wt% CNP on gel strength of commercial invert emulsion drilling

fluids at 10 s (A, B) and 10 min (C, D). ................................................................................ 54

Figure 21: Impact of 0.5 wt% CNP on volumetric composition of virgin (A) and recycled (B)

commercial invert emulsion drilling fluids. .......................................................................... 55

Figure 22: Impact of 0.5 wt% CNP on HPHT fluid loss (A, B) and filter cake thickness (C,

D) of commercial invert emulsion drilling fluids at 80 °C and 500 psi. ............................... 56

Figure 23: HPHT fluid loss as a function of measured depth in the control and test wells at 80

°C and 500 psi. ...................................................................................................................... 60

Figure 24: Average HPHT fluid loss in the control and test wells at 80 °C and 500 psi. ............ 61

Figure 25: HPHT fluid loss and calculated concentration of carrier as a function of measured

depth in the test wells. ........................................................................................................... 63

Figure 26: Cumulative mud losses as a function of measured depth in the control and test

wells. ..................................................................................................................................... 64

Figure 27: Final cumulative mud losses at TD in the control and test wells. ............................... 65

Figure 28: Mud losses per 100 m drilled as a function of measured depth in the control and

test wells. ............................................................................................................................... 66

Figure 29: Average mud losses per 100 m drilled at TD in the control and test wells. ................ 66

Figure 30: Correlation between HPHT fluid loss and instantaneous mud losses in the test

wells. ..................................................................................................................................... 67

Page 11: A Field Application of Nanoparticles For Improved Downhole

x

List of Symbols, Abbreviations and Nomenclature

Symbol Definition

𝐴 Cross-section filtration area (cm2)

API American Petroleum Institute

BHA Bottom-hole assembly

BOP Blowout preventer

BTEX Benzene, toluene, ethylbenzene and xylenes

CMC Critical micelle concentration

CNP Calcium carbonate nanoparticles

𝐷𝑖 Initial depth of displacement to invert (m)

𝐷𝑛 Current measured depth (m)

DLS Dynamic Light Scattering

EDX Energy-Dispersive X-ray Spectroscopy

ES Electrical stability (V)

ℎ𝑐 Filter cake thickness (cm) ℎ𝑓 Formation thickness (cm)

HPHT High pressure, high temperature

IFT Interfacial tension

𝑘𝑐 Permeability of the filter cake (D)

𝑘𝑓 Permeability of the formation (D)

LCM Lost circulation materials

LPLT Low pressure, low temperature

LSD Legal sub-division

MD Measured depth (m)

MW Mud weight (kg/m3)

NPs Nanoparticles

OBM Oil based mud

OWR Oil to water ratio

∆𝑃 Pressure differential (atm)

PDI Polydispersity index

𝑃𝑐 Pressure at the inner boundary of the cake (atm)

𝑃𝑓 Formation pressure (atm)

PV Plastic viscosity (cP)

PVT Pit volume totalizer

𝑃𝑤 Wellbore pressure (atm)

𝑞𝐴𝑃𝐼 API fluid loss (mL/s)

𝑞𝑐 Filtrate flux during transitional period (mL/s)

𝑞𝑠 Rate of spurt loss (mL/s)

𝑟𝑐 Inner filter cake radius (cm)

𝑟𝑒 External formation radius (cm)

ROP Rate of penetration (m/hr)

𝑟𝑤 Wellbore radius (cm)

SG Specific gravity

SEM Scanning Electron Microscopy

𝑉𝐶𝐿 Cumulative mud losses (m3)

Page 12: A Field Application of Nanoparticles For Improved Downhole

xi

𝑉𝐿100 Mud losses per 100 m drilled (m3/100 m)

𝑉𝑡 Volume of total liquid fraction (mL)

𝑉𝑤 Volume of aqueous fraction (mL)

WBM Water based mud

YP Yield point (lb/100 ft2)

Greek Symbols Definition

𝜃600 Dial reading at 600 rpm

𝜃300 Dial reading at 300 rpm

𝜇 Mud viscosity (cP)

𝜇𝑝 Plastic viscosity (cP)

𝜏𝑦 Yield point (lb/100 ft2)

Page 13: A Field Application of Nanoparticles For Improved Downhole

1

CHAPTER ONE: INTRODUCTION

1.1 Functions and properties of drilling fluids

Drilling fluids play a crucial role in oil and gas well drilling operations and provide several key

functions. For instance, drilling fluids are used to suspend and carry cuttings to the surface, provide

hydrostatic pressure and support the wellbore, cool and lubricate the drilling assembly, form a cake

sealing off porous and permeable formations, and transmit signals between downhole equipment

and the surface, to name a few (ASME, 2005). As a result, drilling fluids are complex systems,

often composed of a large number of components, each providing a specific functionality or

property. Moreover, composition and properties of drilling fluids constantly change throughout a

drilling operation. This renders reproducible laboratory analysis challenging, and so only the

samples obtained from the same batch can be compared directly. When developing a mud additive

aimed to improve one property of a drilling fluid, it is important to consider and mitigate its

possible detrimental effects on the other basic properties. Therefore, this section discusses several

critical properties that were the focus of experimental work presented herein, while the standard

lab methods for their evaluation are outlined in Chapter Two.

a) Mud weight. In overbalanced or managed pressure drilling, mud weight is carefully

selected, so that the corresponding hydrostatic pressure exceeds an anticipated formation

pressure. This provides wellbore stability and prevents formation fluids from entering a

wellbore (known as a kick) or reaching the surface and causing a blowout. Conversely, if

a hydrostatic pressure exceeds the formation breakdown pressure, induced fractures can

lead to large volumes of mud losses, environmental contamination, formation damage, and

borehole collapse (Chin, 1995). Since formation pressure gradient increases with depth,

mud weight has to be monitored regularly and maintained within a specific, optimum

Page 14: A Field Application of Nanoparticles For Improved Downhole

2

range. Mud weight can be increased by adding an appropriate amount of weighting

material, such as a barite mineral (barium sulfate, 4.2 SG). Conversely, it can be decreased

either by solids removal or via volumetric dilution with a less dense fluid. American

Petroleum Institute (API) set a standard practice for measuring mud weight in a lab or at a

drilling site using a special mud balance (API, 2014). Mud weight of drilling fluids

investigated in this work was in the range of 900–1150 kg/m3.

b) Rheology and gel strength. Most drilling fluids exhibit non-Newtonian, shear-thinning

behaviour, which can be described by a Bingham plastic rheological model (Agarwal et

al., 2009; Jihua & Sui, 2011). Viscosity of drilling fluid is a very important parameter: it

has to be sufficient to support drilled solids and prevent sagging, yet should not exceed

pumping capabilities of a rig equipment, so that desirable flow regimes in a drilling string

and annular space could develop. Viscosity of drilling fluids also affects their filtration

rates, which will be described in more detail in the next section. An industry-standard

viscometer was used to measure rheology of drilling samples in the lab in order to

determine an impact of nanoparticles addition. Another important feature of drilling fluids

is gel strength. It allows the fluid to thicken when circulation is stopped, such as when

making a connection or during a trip in/out, which suspends solids and prevents their

settling due to gravity.

c) Volume composition. A relative ratio of solid and liquid components in a drilling fluid is

an important parameter that affects final mud properties and is maintained within a range

specified in a mud program. Generally, increase in a volume concentration of suspended

solids has a negative impact on a drilling operation and results in a reduced rate of

penetration (ROP), increased viscosity and mud weight, increased wear on the equipment

Page 15: A Field Application of Nanoparticles For Improved Downhole

3

and formation damage (ASME, 2005). A variety of solids control equipment exists that

help to manage solids content, of which shale shakers and centrifuges are the most common

(ASME, 2005). A typical range of solids found in most drilling fluids is 1–15% vol/vol.

Another important parameter, which applies to oil-based drilling fluids (OBM) that contain

both an oil and an aqueous phase, is the oil to water ratio (OWR). Volume composition of

drilling fluids is measured using a retort analysis, which employs fractional distillation at

high temperature to separate major components of a system.

d) Filtration control. The afore-mentioned overbalanced conditions cause near-wellbore

invasion of drilling mud filtrate and fine particles into permeable, porous or fractured

formations. This leads to reduced return permeability and is known as formation damage.

Depending on the pressure gradient, the nature of wellbore and formation fluids, and the

lithology, very large volumes of drilling fluid can be lost. When fluid loss occurs in the

pay zone, it can lead to significant declines in production. This is partially circumvented

using filtration control agents or lost circulation materials (LCM) that bridge over the pore

or fracture opening, form a filter cake, and reduce the flux of drilling fluid filtrate into the

formation. A wide variety of fibrous and granular materials have been used as LCM to

reduce mud loss, including nut shells, cellophane flakes, crushed rubber, micronized

marble, polymers, bitumen, calcium carbonate, and others (Caenn et al., 2011). An API

fluid loss test is a standard method for performance evaluation of filtration properties of

drilling fluids. Therefore, it served as a primary indicator in development of an optimum

formulation containing nanoparticles.

e) Electrical stability. Electrical stability (ES) is used to quantify stability of W/O emulsions

based on electrical resistance of a liquid medium between two electrodes. A ramped

Page 16: A Field Application of Nanoparticles For Improved Downhole

4

voltage is applied until the current connects the two electrodes. Since an oil phase is non-

conductive, electrical stability of an emulsion is a function of a number, size, and ionic

strength of the dispersed aqueous droplets. In the case of OBM, solid content of an oil

phase can also have an impact on the electrical stability, especially if a high concentration

of conductive particles is present (Growcock et al., 1994). As a result, a decrease in the

electrical stability in an invert emulsion generally indicates an increase in the average size

of water droplets that more readily conduct current than the smaller and better dispersed

ones. A typical electrical stability tester used in the drilling fluids industry is tuned so that

a non-conductive, pure base oil gives the maximum reading of 1999 V, whereas fresh water

gives a reading of 2–5 V. The higher the value, the better the stability of a W/O emulsion,

therefore effects of different mud additives can be investigated.

1.2 Types of drilling fluids

Three major types of drilling fluids can be identified based on their primary continuous phase,

which can be either water, oil, or gas (Caenn et al., 2011). Water-based muds (WBM) and OBM

exhibit different properties, and are selected based on a drilling program, geology, environmental

concerns, and production economics. Gas-based drilling was not consider in this work and

therefore is outside the scope of this discussion.

WBM are based either on fresh water, sea water, or a brine (Caenn et al., 2011). Other

components depend on a particular application and drilling requirements, and may include LCM,

lubricants, polymers, flocculants, dispersants, surfactants, shale stabilizers, corrosion inhibitors,

and others (Caenn et al., 2011). While WBM offer certain advantages compared to OBM, such as

lower cost and minimal environmental impact, they suffer from a number of significant drawbacks.

Page 17: A Field Application of Nanoparticles For Improved Downhole

5

For instance, water has intrinsically greater coefficient of friction than oil, resulting in increased

torque and drag experienced by a drilling string. Moreover, WBM exhibit poor stability at high

reservoir pressures and temperatures, which limits their performance in deep wells or locations

with high a geothermal gradient. Another issue associated with WBM is swelling of shales due to

water adsorption and ionic exchange. This can lead to wellbore instability, stuck pipe, ballooning,

washout, and other undesirable effects (Sensoy et al., 2009).

OBM are typically based on invert emulsions formed by a continuous oil phase and a

dispersed aqueous phase, although some drilling fluids may use base oil alone. On the one hand,

OBM do not suffer from the same problems associated with WBM; but on the other hand, they

have high cost ($1000–$2000/m3 of base oil as of 2015) and are considerably toxic to the

environment (Caenn et al., 2011). Since the upper wellbore sections often contain unconsolidated

and permeable formations, large volumes of mud can be lost, which is very costly and has

disastrous consequences on the environment. As a result, a common practice in the industry is to

drill and complete a surface hole using WBM, followed by OBM for the remainder of a well. A

surface casing is usually set and cemented at a depth of approximately 500–600 m, which serves

to isolate wellbore fluids, prevent their contact with a formation, and protect underground aquifers

from contamination with oil based filtrate. Subsequently, WBM is displaced to invert, a cement

plug is drilled out, and an intermediate well section begins.

1.3 Mud circulation system

A schematic of a typical drilling fluid circulation system is shown in Figure 1. At the surface, 50–

60 m3 of drilling fluid are continuously circulated between several in-line tanks, known as active

tanks. From there, displacement pumps drive the mud up the standpipe and down the drill string,

Page 18: A Field Application of Nanoparticles For Improved Downhole

6

where it exits through high-shear jets of a drill bit. Drilling fluid then travels upward in the annular

space carrying cuttings to the surface. Mud return line passes through blowout preventers (BOPs)

and is directed onto vibrating mesh screens of shale shakers that separate drilling fluid from

cuttings. Wet cuttings are discarded, while mud and suspended solids small enough to pass through

the screens are collected in a settling tank below. Also known as a sand trap, it helps to remove

suspended solids due to gravity segregation. Finally, depending on the amount of desirable fine

solids, rheology, and density, drilling fluid is either redirected to rotor centrifuges first or returned

to active tanks, and the circulation cycle repeats. As the hole volume increases, more drilling fluid

is required to maintain the required volume in active tanks. Volume is increased with fresh mud

that is either mixed on-site using a pre-mix tank or delivered and stored in a rig tank farm. A typical

pre-mix tank has a capacity of 20 m3 and is used to condition a fresh batch of drilling fluid prior

to bleeding it into the active system. A pre-mix tank was also utilized during the field testing of

nanoparticle-based drilling fluids, as described in the experimental section.

Figure 1: A schematic of a rig mud circulation system.

Page 19: A Field Application of Nanoparticles For Improved Downhole

7

1.4 Emulsions and surfactants

Emulsions are thermodynamically unstable mixtures of two or more immiscible fluids that form a

dispersed phase and a continuous phase (Rosen & Kunjappu, 2012). Based on a volume ratio

between phases, emulsions are traditionally classified as either oil-in-water (O/W, direct) or water-

in-oil (W/O, invert). A common example of an O/W emulsion is dairy milk, while OBM are

typically based on W/O emulsions.

Emulsions are formed when an appropriate mixture of fluids, such as water and oil, is

subjected to a high amount of shear. Shear can be in the form of shaking, blending, jet shear, or

sonication (Rosen & Kunjappu, 2012). High interfacial tension (IFT) between immiscible fluids

impacts the size and stability of the dispersed droplets. Smaller droplets have a higher surface

energy than the larger ones, and thus an emulsion is thermodynamically driven toward a lower

total energy. In a static system, this process can be observed as coalescence and growth of the

dispersed pools, which are driven by the Brownian motion and gravity segregation (Boyd et al.,

1972). Eventually, this deterioration of an emulsion resolves in a phase separation and loss of

functionality. The primary mechanism for destabilization of W/O emulsions is a spontaneous

process known as the Oswald ripening, which also affects invert emulsion drilling fluids subjected

to prolonged static conditions, such as during storage or rig down time (Voorhees, 1985).

Stability of emulsions can be greatly improved using a special class of aphiphilic

compounds, also known as surfactants or emulsifiers, which are partially soluble in both phases.

Surfactants contain a hydrophobic tail, typically composed of a hydrocarbon chain, and a

hydrophilic head, which is either a highly polar or an ionic functional group. Solubility of an

emulsifier in water decreases with increasing chain length of a hydrophobic tail, while its solubility

in organic phase increases. When a surfactant is added to a binary mixture of water and oil, its

Page 20: A Field Application of Nanoparticles For Improved Downhole

8

molecules orient along the interface such that its hydrophilic head is in the aqueous phase and its

hydrophobic tail is in the oil phase. This lowers the IFT and surface energy between the phases. A

free energy of the surface is further reduced with increasing concentration of a surfactant, until the

critical micelle concentration (CMC) is reached. The CMC of a surfactant is defined as a bulk

concentration of surfactant, at which its molecules begin to aggregate into micelles. It varies for

each type of a surfactant, depending on temperature, pressure, and the nature of immiscible fluids,

and can be either determined experimentally or obtained from the literature (Mukerjee & Mysels,

1971). Further increase in the concentration of emulsifier above its CMC results in an increased

number of micelles. When shear is applied to such mixture, surfactant molecules dissolve in the

dispersed phase and form a monolayer surrounding the pools. Very small droplets can be formed

due to lower IFT, and their coalescence and growth are stabilized via electrostatic repulsion or

steric hindrance effects of a surfactant film. In some cases, shorter chained hydrocarbons are added

as co-surfactants that help to improve stability of an emulsion. They insert between their larger

counterparts at the interface of a dispersed phase and provide a steric stabilizing effect. As a result,

surfactant film becomes more ‘rigid’ and dispersed droplets reduce in size, which reduces the

possibility of coalescence and growth of droplets (Oh et al., 1993).

1.5 Invert emulsion drilling fluids

A liquid fraction of invert emulsion drilling fluids is composed of a continuous oil phase and a

dispersed water phase, which is stabilized with surfactants. A typical volume ratio of oil to water

(OWR) in commercial drilling fluids varies between 75/25–95/5. While various additives may be

present depending on a particular system, several most common components of commercial invert

emulsions are identified and summarized in this section.

Page 21: A Field Application of Nanoparticles For Improved Downhole

9

a) Base oil. An oil phase of OBM is most commonly represented by long-chain petroleum

distillates, such as diesel (Cutter-D) or gas oil (Distillate 822). These aliphatic base oils

typically contain low aromatics and BTEX, with majority of chains in the C15+ range.

While Cutter-D and Distillate 822 have low cost and show good performance in a variety

of drilling conditions, they are classified as toxic and contain known carcinogens.

Alternatively, some environmentally sensitive drilling applications utilize invert emulsion

drilling fluids based on mineral or synthetic base oil. A common example is the Amodrill

system based on synthetic C12–C14 α-olefins. Synthetic base oil is usually produced via

oligomerization of monomers, which ensures uniform composition and absence of heavy

metal contaminants. Synthetic or mineral base oils have intrinsically lower density and

viscosity than diesel or gas oil and are typically used in deep water drilling, where

formations often have a narrow mud pressure window.

b) Aqueous phase. An aqueous phase of invert emulsion drilling fluids is typically composed

of a brine of calcium chloride with a concentration varying between 20–40% w/w. A high

concentration of chlorides (>100,000 ppm [Cl–]) is used to provide osmotic pressure

against saline interstitial water. The second most common aqueous component is a

saturated solution of calcium hydroxide.

c) Emulsifiers. Most commonly used emulsifiers in oil-based drilling fluids are divalent soaps

of anionic carboxylic acids of varying chain length, degree of branching, and saturation.

Divalent salts of fatty acids are formed by reaction of carboxylic group with Ca2+ ions,

which are usually provided in the form of hydrated lime, Ca(OH)2. Calcium soaps of fatty

acids are better suited to form reverse micelles in W/O emulsions, as opposed to salts of

alkali metals, which are used to prepare O/W emulsions. Commercial blends of emulsifiers

Page 22: A Field Application of Nanoparticles For Improved Downhole

10

are available as technical-grade mixtures of several fatty acids and secondary emulsifiers,

so the exact composition can deviate from batch to batch and is challenging to determine.

Several types of commercial emulsifiers have been investigated in this work, however the

primary component in all cases was a C16–C18 fatty acid, such as palmitic, oleic, or stearic

acid.

d) Wetting agents. Wetting agents are surfactants of varying structure and composition that

are used to impart oil wettability to suspended solids in OBM. Water-wet solids tend to

aggregate and agglomerate, thus increasing rate of solids settling and sagging.

Furthermore, adsorption of wetting agents by water-sensitive shales results in hydrophobic

rock surface that helps to prevent swelling. Wetting agents are commonly included as part

of commercial blends of primary and secondary emulsifiers, but also are available as

separate products at most rig sites.

e) Hydrated lime. Hydrated lime, slacked lime or lime, are all common names for calcium

hydroxide, which is used to activate anionic emulsifiers in W/O drilling fluids. While

excess lime remains in the suspension and contributes to a total solids fraction, a small

amount is dissolved in the aqueous phase reaching its saturation point (1.73 g/L at standard

conditions). This functions to provide alkalinity to the aqueous phase, where pH above 10–

11 is preferred to improve interaction with anionic surfactants.

f) Organophilic clay. A surface-modified bentonite clay is used in W/O drilling fluids as a

rheology modifier and a filtration control agent. With its surface coated with an

organophilic compound, bentonite readily disperses in oil phase even at low shear and

increases viscosity and gel strength of a mud. In addition, clay platelets serve as bridging

agents and reduce mud filtration.

Page 23: A Field Application of Nanoparticles For Improved Downhole

11

g) Calcium carbonate. Calcium carbonate is the most abundant LCM used in most WBM and

OBM. It is the primary component in limestone rock and typically is produced in industry

by crushing and milling limestone rock, making it an inexpensive and benign material.

Several grades of calcium carbonate are available based on the mean particle size, which

normally ranges between 10 µm–1 mm (Cargnel & Luzardo, 1999). Most drilling

applications utilize blends of several calcium carbonate grades to achieve broad size

distribution in the mud. A long history of application of calcium carbonate in drilling fluids

rendered it a good candidate for nanoparticle research and field implementation, which is

a focus of this contribution.

h) Gilsonite. Gilsonite is a naturally occurring asphaltene hydrocarbon used as a loss control

additive in OBM. Its particle size varies between 150 µm–5 mm, and it is responsible for

a characteristic black-brown colour of invert emulsion drilling fluids.

i) Barite. Barite is a natural mineral from of barium sulfate. It is a dense material (SG 4.2)

and is used as a weighting agent in drilling fluids. Due to limitations of mechanical crushing

used in a large-scale production of barite, particle size distribution is typically broad, which

may increase possibility of sag, especially in lateral wellbore sections or when large

quantities of barite are required to increase mud weight.

j) Graphite. Granular graphite with varying particle size is used in drilling fluids as an LCM

and lubricant additive.

1.6 Mechanism of fluid loss and cross-flow filtration

Dynamic filtration of drilling fluids has been investigated extensively in an attempt to mitigate

mud losses and formation damage in troublesome geological zones, such as porous, permeable,

Page 24: A Field Application of Nanoparticles For Improved Downhole

12

unconsolidated, fractured, vuggy, or depleted formations (Al-Hitti et al., 2005; Bennion et al.,

1997; Chin, 1995). Three main stages of fluid flow can be identified in a theoretical analysis of

dynamic mud filtration, namely spurt loss, cake growth, and cake equilibrium (Jiao & Sharma,

1994).

Spurt loss refers to the flux of filtrate at the early times of filtration, before the filter cake

is formed. It can be approximated using a radial form of Darcy’s law, which is given by the

following flow equation:

𝑞𝑠 =

2𝜋𝑘𝑓ℎ𝑓(𝑃𝑤 − 𝑃𝑓)

𝜇ln(𝑟𝑒𝑟𝑤)

(Eq. 1)

where 𝑞𝑠 is the rate of spurt loss (mL/s); 𝑘𝑓 is the permeability of formation (D); ℎ𝑓 is the formation

thickness (cm); 𝑃𝑤 is the wellbore pressure (atm); 𝑃𝑓 is the formation pressure (atm); 𝜇 is the

viscosity of mud filtrate (cP); 𝑟𝑒 is the external radius of formation (cm); 𝑟𝑤 is the wellbore radius

(cm).

Eq. (1) shows that the dynamic flow rate of filtrate in the absence of a filter cake is directly

proportional to the permeability of the formation and the pressure gradient, and inversely

proportional to the viscosity of the mud. As a result, spurt loss in overbalanced drilling can be very

high, especially in the case of highly permeable formations, leading to deep invasion of mud,

damage, and high losses.

Flow of filtrate in the radial direction gives rise to a hydrodynamic drag force, which carries

suspended particles towards the borehole wall. Depending on the particle/pore size ratio, they

either pass through larger openings or get retained at the wall. Liquid and solid particles that pass

through the opening travel further into the formation, where they are eventually immobilized due

to the combined effect of sedimentation, direct interception, and surface attractive forces (Bezeme

Page 25: A Field Application of Nanoparticles For Improved Downhole

13

& Havenaar, 1966). Larger particles normally deposit first followed by smaller ones, as per the

size exclusion principle, and an internal filter cake begins forming. This can cause severe

permeability impairment and formation damage. Conversely, particles that are approximately one

third the size of the opening and larger, get screened out at the wellbore wall and form the external

filter cake. Filtrate flux through the cake becomes a function of pressure drop across the cake as

well as its thickness and permeability. The transitional flow regime can represented

mathematically by the following equation:

𝑞𝑐 =

2𝜋𝑘𝑐ℎ𝑓(𝑃𝑤 − 𝑃𝑐)

𝜇ln(𝑟𝑤𝑟𝑐)

(Eq. 2)

where 𝑞𝑐 is the filtrate flux during transitional period (mL/s); 𝑘𝑐 is the permeability of the filter

cake (D); 𝑃𝑐 is the pressure at the inner boundary of the filter cake (atm); 𝑟𝑐 is the inner radius of

the filter cake (cm).

As more particles are deposited at its surface, the cake continues to grow, and its

permeability and thickness increase, causing reduction of the filtrate flow rate. Consequently,

pressure drop reduces, and the drag force becomes weaker, so that only smaller and smaller

particles are delivered and deposited at the cake surface, until no particles small enough are present

in the mud. This gives rise to an inhomogeneous distribution of particles throughout the cake,

which in turn affects particle packing, porosity, and permeability. Cake growth rate decreases until

it reaches its equilibrium thickness, at which point a steady state filtrate flow is established. Under

dynamic conditions, circulating mud exerts shear stress and erodes the surface of the cake,

therefore its characteristic thickness is achieved when a normal force keeping particles at its

surface arrives at equilibrium with a tangential force acting to detach them.

Page 26: A Field Application of Nanoparticles For Improved Downhole

14

Jiao and Sharma (1994) proposed a widely accepted mechanism for filter cake build up,

which concluded that the solids deposited at the cake surface experience colloidal interactions with

neighbouring particles. Surfaces forces include van der Waals, electrostatic, structural, and Born

forces (Jiao & Sharma, 1994). As the mean particle size decreases, the surface forces act over

shorter separation distances and become more significant. At very small, submicron or nanometer

separation distances, surface forces become orders of magnitude greater than the hydrodynamic

forces, which themselves act over much longer distances. This leads to irreversible particle

attachment and increases resistance of the cake to erosion.

The aforementioned mechanism suggests that the extent of drilling filtrate invasion and

formation damage is strongly dependent on the particle/opening size ratio and can be drastically

reduced if a cake with low permeability is formed rapidly. Since geology can vary greatly

throughout a given well, typical drilling fluids contain a mixture of loss additives to attain wide

particle size distribution and effectively seal appropriately sized openings. However, commercial

filtration materials range anywhere from several µm to several mm in size, which makes them

unsuccessful creating a filter cake in the cases where openings are in the submicron domain, such

as in shales or induced micro-fractures (Nelson, 2009). Given that enough pressure gradient is

provided, mud filtrate can still invade into such formations and cause a number of problems,

including fracturing, clay swelling, obstructed formation evaluation, borehole instability, washout

or collapse (Zamora et al., 2000). A proposed model of filtration in the presence of conventional

micron-sized LCM and NPs is illustrated in Figure 2 (A and B, respectively).

Page 27: A Field Application of Nanoparticles For Improved Downhole

15

Figure 2: Drilling fluid filtration in the presence of conventional LCM (A) and NPs (B).

Filtration properties of drilling fluids can be evaluated in the lab using an industry-standard

American Petroleum Institute (API) fluid loss test. An API low pressure, low temperature (LPLT)

fluids loss experiment applies to WBM, while OBM are tested at high pressure and temperature

(HPHT) to approximate reservoir conditions. Both tests share the same principle and are based on

static filtration of mud through a specially-hardened filter paper, which substitutes a porous

medium. Filtrate is collected at specified time intervals, and the total volume of filtrate collected

in 30 min is used to quantify performance of drilling fluids. Filtrate flux in this case is given by a

one dimensional form of Darcy’s law (Chelton, 1967):

𝑞𝐴𝑃𝐼 =

𝑘𝑐𝐴∆𝑃

𝜇ℎ𝑐 (Eq. 3)

where 𝑞𝐴𝑃𝐼 is the API fluid loss rate (mL/s); 𝐴 is the cross-sectional filtration area (cm2); ∆𝑃 is the

pressure differential (atm); ℎ𝑐 is the thickness of the filter cake (cm).

Page 28: A Field Application of Nanoparticles For Improved Downhole

16

API fluid loss is typically performed at fixed pressure differential, temperature, and surface

area, so that the volume of filtrate is only a function of viscosity of mud and cake parameters.

While the API fluid loss test is the most accepted practice in the drilling industry, it poorly

represents the actual conditions encountered in dynamic filtration through a permeable rock.

Correlation between lab fluid loss and real downhole losses has long been a subject of debate, but

the general observed trend is that subsurface losses tend to be lower in drilling fluids that provide

low LPLT or HPHT volumes (Villas-Boas et al., 1999). Other less common lab techniques used

to evaluate loss prevention include dynamic filtration, core testing, and permeability plugging (Al-

Riyama & Sharma, 2004).

1.7 Nanoparticles in oil and gas

According to a strict definition, nanoparticles are classified as materials with a size in the range of

1–100 nm (Cao, 2004). However, it should be pointed out that in the context of this work, the term

‘nanoparticles’ is used more loosely to refer to a range of submicron size distributions spanning

between 1–1000 nm.

As a particle becomes smaller, its surface area, which is defined as m2/g, increases

accordingly. This leads to a high surface activity and imparts unique physical, chemical, electrical,

thermal, and reactive properties to nano-scale materials that are typically not exhibited by their

larger counterparts with identical chemical structures. Nanotechnology has seen wide-spread

applications in such diverse industries as medicine, electronics, fuel cells, food, and various

consumer products (Cao, 2004). In recent years, a significant research effort focused on potential

applications of nanomaterials in the oil and gas industry. Some of the areas that showed useful

benefits of NPs include heavy oil upgrading (Nassar et al., 2011), enhanced oil recovery (Haroun

Page 29: A Field Application of Nanoparticles For Improved Downhole

17

et al., 2012; Skauge et al., 2010), hydraulic fracturing (Huang et al., 2010), cementing operations

(Patil & Deshpande, 2012), and drilling, drill-in, completion, stimulation, and workover fluids

(Amanullah & Al-Tahini, 2009; Zakaria et al., 2012). Applications of nanomaterials in drilling

fluids are reviewed in more detail in the paragraphs that follow.

Majority of wellbore instability problems associated with WBM are caused by invasion of

water filtrate into shale formations, which in turn comprise 75% of all the lithologies drilled

(Sensoy et al., 2009). Often, the only alternative is to drill with OBM instead; however, this

drastically increases drilling cost and can be damaging to the environment. As a result, a number

of recent studies investigated the use of nanoparticles to increase shale stability. For instance,

Sensoy et al. (2009) used a pressure transmission technique to show that water invasion in shale

samples was reduced in the presence of commercial silica nanoparticles. Water-based mud samples

were prepared using two sizes of ex-situ particles, 5 nm and 20 nm, while their concentration

varied between 5–40 wt%. Experimental results showed that nanoparticles reduced filtrate

invasion in Atoka shale by 98% compared to pure sea water. Recycled drilling fluids also benefited

from the addition of NPs, where fluid invasion was reduced by 16–72% in Atoka shales and by

17–27% in the Gulf of Mexico shale. Scanning Electron Microscopy of shale samples revealed

that nanoparticles and their agglomerates were effective plugging a range of pore throat sizes,

which lead to a 10-fold reduction in permeability. However, noticeable performance was only

achieved at concentrations of NPs above 10 wt%.

Cai et al. (2012) reported similar results in permeability plugging experiments on Atoka

shale samples using seven different commercial nanomaterials with the size between 7–15 nm.

Study of the effect of nanoparticle concentration suggested that fluids containing 10 wt% of NPs

reduced permeability of shale samples by 73–99%.

Page 30: A Field Application of Nanoparticles For Improved Downhole

18

Amanullah et al. (2011) formulated several water-based drilling fluids using three

commercial nanomaterials. It was shown that properties of mud samples were enhanced at low

concentration of NPs (<0.5 wt%) compared to formulations containing conventional macro-sized

additives. In addition to good stability, rheology and gel strength, nano-fluids reduced API spurt

loss and formed a thin mudcake, which is highly desirable. The superior performance of nano-

fluids was attributed to surface interactions that were more dominant on a nano-scale than the

physical forces.

Zakaria et al (2012) investigated the application of nanoparticles for fluid loss reduction in

invert emulsion drilling fluids. Both commercial as well as in-house nanomaterials were tested,

which showed that particles synthesized within the drilling fluid exhibited far superior

performance. Experimental results suggested that in situ NPs were well dispersed and stabilized,

and therefore interacted more favourably with other components in the matrix. While commercial

nanomaterials provided only a negligible LPLT fluid loss reduction of 7%, in-house particles

minimized spurt loss and resulted in 70% lower filtrate volumes.

Nwaoji et al. (2013) introduced a drilling fluid that contained a blend of graphite and in-

house nanoparticles. Hydraulic fracturing experiments were conducted on Roubidoux sandstone

and concrete core samples to investigate potential applications of NPs in wellbore strengthening.

Fluids containing iron- or calcium-based NPs increased fracture pressure resistance of sandstone

by 70% and 36%, respectively. Furthermore, a 25% increase in fracture pressure of impermeable

concrete cores suggested that NPs could provide wellbore strengthening in tight or shale

formations.

Contreras et al. (2014a, 2014b, 2014c) further demonstrated that in-house nanoparticles at

concentrations of 0.5–2.5 wt% reduced HPHT fluid loss in commercial drilling fluids and

Page 31: A Field Application of Nanoparticles For Improved Downhole

19

increased strengthening in sandstone and shale cores. Fracture pressure in sandstone was increased

by up to 65%, while a 30% increase was achieved in shale samples. SEM and EDX analysis of

core samples showed that the induced fractures were sealed with NPs along their entire length,

which suggested a tip screen-out mechanism of wellbore strengthening (Contreras et al., 2014c).

1.8 Emulsion-based synthesis of nanoparticles

There exists a variety of techniques used in the manufacturing of nanoparticles, which are

arbitrarily classified as either dry or wet (Husein & Nassar, 2008). Dry techniques normally

employ mechanical milling to reduce the average particle size (Midoux et al., 1999). Alas, such

approach makes it difficult to control the average size, shape, and properties of the final materials.

Conversely, wet techniques are based on the formation of nanomaterials starting from chemical

precursors and include chemical coprecipitation, sonochemical, electrochemical, reverse micelles,

and sol-gel methods (Husein & Nassar, 2008). Unlike the dry approach, wet techniques can be

tailored to produce nanoparticles with desirable properties (Jolivet et al., 2004). Furthermore,

nanomaterials produced via a wet chemical reaction do not require handling of nano-powders and

hence do not produce potentially hazardous airborne particles.

Chemical coprecipitation of aqueous precursors remains one of the most economic and

versatile processes that could be implemented on a large scale. The method is based on a reaction

between appropriate aqueous solutions that forms an insoluble product, and usually requires a

control mechanism for achieving size distribution in the nanometer domain. In addition to a wide

variety of inorganic nano-materials that can be produced via this approach, their shape, size, and

surface properties can be fine-tuned to meet the required specifications (Parida et al., 2009).

Page 32: A Field Application of Nanoparticles For Improved Downhole

20

Two main stages can be identified in the process of coprecipitation in bulk aqueous phase:

nucleation and growth (Koutsoukos & Kontoyannis, 1984; Taubert et al., 2002). During the

primary nucleation, anionic and cationic species combine, reach local supersaturation, and

precipitate to form an initial crystal. Once formed, seed crystals affect the formation of subsequent

crystals via the process of secondary nucleation. Crystal grows by coming in contact with dissolved

precursors, where it functions as a substrate for the formation of sequential layers. Since smaller

crystals have higher free energy, large particle size is thermodynamically preferred. As a result,

particles have broad size distribution in the micron or even millimeter domain, which is illustrated

in Figure 3 (top).

Size control during coprecipitation in bulk aqueous phase can be achieved by separating

the nucleation and growth processes, which is accomplished either mechanically or chemically.

High shear conditions disrupt crystal growth and increase the number of nucleation sites, which

limits the final particle size. Alas, such high shear mixing is difficult to attain on a large scale, as

was learned in the process of this work. Chemical methods of size control may include surfactants,

polymers or other materials capable of interacting with a crystal face and preventing its growth

(Taubert et al., 2002).

Page 33: A Field Application of Nanoparticles For Improved Downhole

21

Figure 3: A schematic of chemical coprecipitation using bulk aqueous phase (top) and a reverse

micelle approach (bottom).

An alternative to a bulk-phase reaction is known as a reverse micelle or microemulsion

approach (Husein & Nassar, 2007a, 2007b, 2008). This method utilizes invert emulsions, in which

micelles act as individual reactors during coprecipitation. A schematic of a typical process is

provided in the bottom of Figure 3 and can be described as follows. First, two separate invert

emulsions are prepared for each of the precursors using a stoichiometric mixture of oil phase,

surfactants, and aqueous solutions. Next, the invert emulsions are mixed together and shear is

applied. Dispersed water droplets collide, coalesce, exchange solute molecules, and break up

again, until the precursors are fully consumed. Since the amount of available ions, that otherwise

would contribute to crystal growth, is limited by the concentration in each individual droplet, the

average particle size tends to be restricted to a nanometer domain (Husein & Nassar, 2008).

Furthermore, nanoparticles either remain dispersed within the individual pools or interact with a

hydrophilic group of emulsifiers and cross the monolayer into the oil phase, where they are

Page 34: A Field Application of Nanoparticles For Improved Downhole

22

stabilized due to a repulsive effect of a surfactant. This prevents collision and agglomeration of

particles and ensures a narrow size distribution, which often cannot be achieved in a bulk phase

process.

Similar principle, albeit a few modifications, was adopted for an in situ synthesis of

calcium carbonate nanoparticles in invert emulsion drilling fluids. Instead of two separate

emulsions, a single drilling fluid system was utilized. The first aqueous precursor was added to an

invert, where it became emulsified and introduced solute molecules to the already existing water

droplets. Subsequently, the second aqueous precursor was titrated slowly into the system under

continuous mixing. Water droplets containing a counterion were then formed, and a

coprecipitation reaction proceeded according to the two-emulsion mechanism described above.

1.9 Research objectives

Work by Zakaria et al. (2012) served as a starting point of this project. On the one hand, it was

demonstrated that not only invert emulsion drilling fluids can be utilized to synthesize in situ NPs,

but also that such samples exhibited much better filtration properties than those containing ex-situ,

commercial NPs. On the other hand, the process targeted low concentration of NPs in order not to

compromise the stability of the resultant dispersion and was limited to laboratory-scale

preparation. Moreover, it is also important to prove the concept by testing a broad variety of

drilling fluid systems, both virgin and recycled. To summarize, this project focused on the

following key research objectives:

1) Modify the existing technique for the in situ synthesis of calcium carbonate NPs in invert

emulsion drilling fluids in order to ensure a scalable, safe, and cost-effective process.

Page 35: A Field Application of Nanoparticles For Improved Downhole

23

2) Expand the variety of tested drilling fluids in order to prove the concept. In addition to fluid

loss, investigate the impact of NPs on mud weight, OWR, solid content, ES, and rheology

in virgin and recycled mud systems.

3) Develop a scalable method for manufacturing of NPs based on a concentrated carrier

emulsion. Optimize the formulation to achieve a stable emulsion at a high concentration of

NPs and compare performance of this dispersion delivery mechanism to that of the direct,

in situ synthesis.

4) Establish a baseline performance of a final lab formulation using several common invert

systems and replicate the process on a large scale. Conduct a full-scale field test of the final

formulation in a live well. Analyze the data and determine potential impact on losses and

a drilling operation as a whole.

5) Develop a theoretical framework to interpret the results and investigate the applicability of

carrier emulsion approach in synthesis of other types of NPs. Develop standard operating

procedures and provide recommendations for future work.

Page 36: A Field Application of Nanoparticles For Improved Downhole

24

CHAPTER TWO: MATERIALS AND METHODS

2.1 Chemicals and precursors

All chemicals were used as received, unless specified otherwise. Calcium chloride anhydrite

(technical grade, 96-98%) and potassium carbonate anhydrite (technical grade, 95%) were

obtained from Univar Canada (Calgary, Alberta) and used as precursors in the lab bench synthesis

of calcium carbonate NPs. Salts were dissolved in deionized water to prepare 45% w/w and 50%

w/w stock solutions, respectively. Dissolution reaction is exothermic is both cases, so proper

precautions must be taken. Stock solutions were stored at room temperature without visible

recrystallization and used as required.

A large-scale synthesis of calcium carbonate was based on a 50% w/w solution of K2CO3

and 45% w/w solution of CaCl2, both supplied by Univar Canada (Calgary, Alberta). Solutions

were received in 1 m3 totes and stored at a mixing facility where product manufacturing took place,

until required. A Distillate 822 base oil, a premium blend of emulsifiers and wetting agents,

Bentone 150 organophilic clay, and hydrated lime were obtained from Gibson Energy (Sexsmith,

Alberta, Canada).

2.2 Drilling fluids sample preparation

Three virgin and three recycled samples of drilling fluids were tested in the lab, as summarized in

Table 2-1 below. In order to eliminate effect of shear on the measured properties of the samples, a

standard operating procedure for mixing and handling was developed based on the API

Recommended Practices (API, 2014) and International Organization for Standardization

procedures (ISO 10416:2008, 2014). Drilling fluids were first well mixed either via vigorous

shaking or using a stationary dispersator operating at 3,000–4,000 rpm (FANN, USA) to ensure

Page 37: A Field Application of Nanoparticles For Improved Downhole

25

homogenous composition. Subsequently, they were subjected to high shear for 30 min. Depending

on the sample volume, either a Hamilton Beach HMD-200 mixer operating at 7,500 rpm or a

Waring Lab Blender (Model #6012G) operating at 6,500 rpm was used. Finally, fresh samples

were allowed to age and develop at ambient conditions for 10–15 min before being tested. A

typical sample volume ranged between 200–600 mL, and minimum of three replicates per

experiment were sheared separately to reduce experimental errors. Control samples and

formulations containing in situ NPs were obtained from same batches of drilling fluids to ensure

reproducibility.

Table 2-1: A summary of commercial drilling fluid lab samples.

Name Supplier Base Oil Virgin/Recycled Mud Weight (kg/m3) O/W Ratio

Cutter-D Bri-Chem Diesel Virgin 1030 90/10

Escaid 110 DSCo Mineral Virgin 913 80/20

Diesel OBM DSCo Diesel Virgin 975 78/22

Cutter-D Blackstone Diesel Recycled 1070 85/15

Distillate

822 Blackstone Diesel Recycled 1090 85/15

Megadrill MI Swaco Diesel Recycled 1150 85/15

2.3 Drilling fluids testing

2.3.1 Mud weight

Mud weight was measured at 22 °C and standard pressure using a FANN Model 140 mud balance

(Part No. 206768, FANN, USA), shown in Figure 4. Three measurements were collected per

sample following a high-shear mixing protocol, as described above. Prior to taking a reading, mud

holder was tapped with a metal cap to release bubbles of dissolved gas.

Page 38: A Field Application of Nanoparticles For Improved Downhole

26

Figure 4: A FANN Model 140 mud balance.

2.3.2 Electrical stability

Electrical stability of invert emulsion drilling fluids was measured with an OFITE electrical

stability tester (Model #131-50, OFI, USA), which is shown in Figure 5. The meter was calibrated

on a regular basis using standards. Sample of drilling fluid was first subjected to 30 min shear and

then transferred into a FANN heating cup (Part No. 101558383, FANN, USA) set at 50 °C.

Temperature was allowed to equilibrate over a 10 min period, after which a sample was rapidly

stirred for 2–3 sec using a probe, and the reading was taken.

Figure 5: An OFITE electrical stability tester (Model #131-50).

Page 39: A Field Application of Nanoparticles For Improved Downhole

27

2.3.3 Rheology and gel strength

A Fann Model 35 viscometer (FANN, USA) was used to measure rheology of drilling fluids

samples, as shown in Figure 6. Similarly to the procedure for electrical stability described above,

samples were sheared for 30 min and transferred to a heating cup. Heating cup was raised to cover

the bob and the viscometer was set to 600 rpm for 10 min. Subsequently, the readings were taken

at 600, 300, 200, 100, 6, and 3 rpm. Finally, gel strength was measured at 10 sec and 10 min. Gel

strength is an important property of drilling fluids, which allows them to thicken and support solids

during stopped circulation. It is defined as the peak meter reading achieved when a bob is rotated

at 3 rpm following an appropriate static period. Readings at 600 and 300 were used to calculate

yield point and plastic viscosity per the following equations:

𝜇𝑝 =𝜃600 − 𝜃300 (Eq. 4)

𝜏𝑦 =𝜃300 − 𝜇𝑝 (Eq. 5)

where 𝜇𝑝 is plastic viscosity (cP); 𝜏𝑦 is yield point (lb/100 ft2); 𝜃300 and 𝜃600 are dial readings at

300 and 600 rpm, respectively.

Figure 6: A Fann Model 35 viscometer.

Page 40: A Field Application of Nanoparticles For Improved Downhole

28

2.3.4 HPHT fluid loss test

HPHT experiments were carried out using a 175 mL single-capped OFITE HPHT filter press with

a 40 cm2 filtration area (Model #170-00, OFI, USA), shown in Figure 7. Samples were sheared

and ~160 mL were loaded into a mud cell. Cell was then assembled and placed inside a heating

jacket, where it was allowed to reach the target temperature over a 30 min period. Temperature

varied between 80–120 °C to match the testing conditions used by mud companies. During the

first 30 min, 100 psi of pressure were applied by means of compressed CO2 bulbs for 30 min to

suppress vapour pressure build up. Subsequently, pressure differential was increased to 500 psi

and the bottom valve stem was opened to initiate filtration. Volume of filtrate was recorded after

30 min and reported as a multiple of 2 to convert the filtration area to the API standard. Cell was

then emptied and the thickness of a mud cake was measured using a digital calliper with accuracy

of ±0.1 mm. Three readings were taken at different sites of a cake and reported as an average.

Figure 7: An OFITE HPHT filter press (Model #170-00).

Page 41: A Field Application of Nanoparticles For Improved Downhole

29

2.3.5 Retort analysis

Retort analysis of mud composition was performed using an OFITE 50 mL retort kit (Part #165-

14), as depicted in Figure 8. Sample cup was filled with a freshly sheared invert in several portions,

followed by gentle tapping to release dissolved gas, until the fluid level reached the top of the cap.

Fine steel wool was packed into the top part of the holder to prevent non-volatile components from

entering the condenser. Liquid fraction was collected in a 50 mL graduated cylinder, and the

experiment was stopped when the volume remained constant for 10 min. Total volume in the

cylinder, as well as the volume of the bottom aqueous phase were recorded and used to calculate

volumetric composition of the mud as follows:

𝑆𝑜𝑙𝑖𝑑𝑠(𝑣𝑜𝑙%) =

50 − 𝑉𝑡50

× 100% (Eq. 6)

𝑂𝑊𝑅 =

𝑉𝑡 − 𝑉𝑤𝑉𝑡

× 100 ∶ 𝑉𝑤𝑉𝑡

× 100 (Eq. 7)

This approach provides ‘uncorrected’ volumetric ratios as it does not account for the

concentration of chlorides and other ions in the aqueous phase, which remain in the solid phase

after evaporation of liquid components. However, the addition of the NPs results in only negligible

increase in chlorides, so this method was deemed appropriate for relative comparison between

control samples and samples with NPs.

Page 42: A Field Application of Nanoparticles For Improved Downhole

30

Figure 8: An OFITE 50 mL retort kit (Model #165-14).

2.4 In situ synthesis of NPs

In situ synthesis of calcium carbonate NPs, both bench- and field-scale, was based on a reverse

micelle method described previously. Bench scale synthesis was carried out using a 300 mL

sample volume and a Hamilton Beach mixer equipped with a wave spindle and operating at 7,500

rpm. The reaction between aqueous precursors proceeds according to the following equation:

𝐶𝑎𝐶𝑙2(𝑎𝑞) + 𝐾2𝐶𝑂3(𝑎𝑞) → 𝐶𝑎𝐶𝑂3(𝑠) + 2𝐾𝐶𝑙(𝑠)

In situ synthesis was carried out as per the following procedure. First, a target volume of

mud with a known density was determined and used to calculate the mass of calcium carbonate

required to achieve a desirable concentration, which in the case of this work was 0.5 wt%. Next,

reaction stoichiometry and molecular weights of reactants and products were used to calculate the

mass of the precursors. The latter was then converted to the mass of a corresponding aqueous

solution with known concentration. Finally, the mass was divided by a density of aqueous solutions

(measured or theoretical) to obtain the required volume of each of the precursors. Appropriate

volumes of precursors were drawn into 10 mL plastic syringes and added slowly, drop wise into a

Page 43: A Field Application of Nanoparticles For Improved Downhole

31

shearing invert sample. Calcium chloride was added first, which was followed by a 10 min mixing.

After sufficient time was given for droplets of calcium chloride solution to reduce in size and

interact with emulsifiers, potassium carbonate was added slowly to the system. Sample was

sheared for additional 30 min to ensure formation of calcium carbonate NPs. A slight color change

and temperature increase indicated that the coprecipitation reaction took place.

2.5 Carrier emulsion synthesis

The aforementioned in situ formation of NPs typically requires high shear and good micro-mixing

to produce submicron particles, which is difficult to realize on a large scale. Instead, a custom

invert emulsion with a high concentration of calcium carbonate NPs was produced and used as a

carrier to deliver NPs to a compatible drilling fluid via volume dilution. Both bench- and field-

scale methods were modeled after the in situ preparation and employed the same aqueous

precursors to precipitate calcium carbonate NPs. To ensure the possibility of scale up, a

formulation was based on commonly used and commercially available mud products, and its basic

properties were designed to closely match those of typical invert emulsion drilling fluids. Cost

effectiveness of a large-scale production via this approach depends on the concentration of NPs

that can be achieved in a carrier emulsion. On the one hand, a higher concentration would result

in a lower volume of carrier required to achieve a target concentration of NPs on-site. On the other

hand, since this method utilizes aqueous precursors, there is a limit on how much water an invert

emulsion system can accommodate before it separates. Furthermore, high concentration of calcium

carbonate would inevitably affect the nucleation and growth processes, where abundance of

suspended crystals would increase rates of aggregation and agglomeration. Long-term observation

has shown that a 5 wt% concentration of NPs provided the optimum emulsion quality, which was

Page 44: A Field Application of Nanoparticles For Improved Downhole

32

stable after 1 week of static aging at ambient conditions. Various combinations of components

were tested until a formulation with the desirable properties was achieved. A typical formulation

of a carrier emulsion with 5 wt% concentration of calcium carbonate NPs is provided Table 2-2.

Table 2-2: A typical formulation of a carrier emulsion with 5 wt% of CNP.

Component Concentration

Base Oil 740–760 L/m3

Emulsifiers 35–45 L/m3

Hydrated Lime 27–30 kg/m3

CaCl2 (aq) 45% w/w 105–115 L/m3

K2CO3 (aq) 50% w/w 90–100 L/m3

Bentone 150 7.5–8.0 kg/m3

2.5.1 Bench-scale synthesis of a carrier emulsion

Lab-scale samples with a total volume of 300 mL were prepared using a Hamilton Beach HMD-

200 mixer. First, base oil was added to a mixing cup and sheared for several minutes. Next, a blend

of emulsifiers was added to the base oil and allowed to mix for additional 5–10 min. Subsequently,

an appropriate amount of hydrated lime powder was slowly added to a cup and allowed to react

with anionic surfactants over a 15 min period of time. Once sufficient time was given, aqueous

calcium chloride was added to the cup drop-wise while mixing. Shearing continued for 30 min to

allow formation of invert emulsion with a small droplet size. Aqueous potassium carbonate was

then added drop-wise. The rate of addition was adjusted to control exothermic effect of a

coprecipitation reaction, such that the temperature remained below 60–70 °C. Following the

addition of the second precursor, invert emulsion was mixed for a total of 30 min. Within 10 min,

Page 45: A Field Application of Nanoparticles For Improved Downhole

33

its colour changed from dark brown to grey-white, which indicated formation of white calcium

carbonate precipitate. Bentone 150 was the last component added to the system and was used to

impart rheological properties and increase viscosity of an emulsion. Invert emulsions in the

absence of clays were not able to support high content of solids, which increased rates of settling

and separation of a suspension. The final product was mixed for additional 30 min before testing.

2.5.2 Field-scale synthesis of a carrier emulsion

Carrier emulsion for field testing was produced on a 20 m3 scale at a Gibson Energy toll mixing

facility located in Sexsmith, Alberta, Canada. Large-scale process utilized the same reagents,

concentrations, sequence and rate of addition, and mixing duration as in the bench-scale process

described previously. The facility was equipped to produce large batches of various OBM systems

and could accommodate volumes up to 75 m3 per batch. The primary mixing skid consisted of

three in-line tanks with 25 m3 capacity. Rather than using high-energy mechanical mixing to

achieve good shear and produce an emulsion, shear jet nozzles and a 150 HP centrifugal pump

were employed. The principle is based on injecting fluid under high pressure through a specially

designed constriction, thus creating a cavity on the other side of a jet. Significant pressure

differential causes the cavity to implode, which sends propagating shock waves through the

medium and achieves high-shear mixing.

The synthesis started by transferring an appropriate volume of Distillate 822 base oil to a

mixing skid, which was followed by emulsifiers. Hydrated lime was then added through a solids

hopper, which was also equipped with a shear jet to ensure good mixing and dispersion of solid

additives. The mixture was circulated for 1 h, and calcium chloride brine was slowly added to the

system through a mixing line at a rate of 40 L/min. Circulation continued for 1 h to ensure

Page 46: A Field Application of Nanoparticles For Improved Downhole

34

formation of an invert emulsion. Subsequently, potassium carbonate brine was introduced in a

similar fashion at a rate of 30 L/min. Temperature of a flow line was monitored at regular intervals

to avoid excessive heat generation. Emulsion was mixed for additional 1 h to ensure completion

of the coprecipitation reaction before adding Bentone 150. Finally, the product was circulated for

1–2 h and transferred to a tanker truck for delivery to a rig site. At the same time, several litres of

carrier were collected for analysis. Mud weight and electrical stability were measured at the mixing

facility, while other properties were evaluated in a lab at a later time.

2.6 Particle characterization

Conventional analytical methods used to characterize nanoparticles are highly sensitive to

concentration and size distribution of solids, as well as the nature and homogeneity of a dispersant.

For instance, particle sizing techniques based on light diffraction (e.g. dynamic light scattering or

laser scattering) rely on a low degree of sample polydispersity and absence of large, agglomerating

or settling particles. As was mentioned earlier, drilling fluids are very complex system and contain

many solid additives, which gives rise to a broad size distribution. As a result, it becomes

extremely challenging to isolate and characterize just one component of the system. In addition,

invert emulsions are multi-component mixtures, therefore an aliquot of a well-mixed drilling fluid

sample contains a continuous oil phase, dispersed water droplets, and micelles of surfactant

molecules. This further hinders accurate analysis of particle size distribution using light scattering

techniques. A Malvern Zetasizer Nano-ZS (Malvern Instruments, UK) was used to analyze size of

calcium carbonate NPs produced in situ, but the data quality was rather low.

Similarly, Electron Scanning Microscopy (SEM) is routinely used to observe

nanoparticles, but high resolution is necessary to distinguish sub-micron particles, which is not

Page 47: A Field Application of Nanoparticles For Improved Downhole

35

attainable with all instruments. Also, SEM chamber operates under vacuum, which precludes the

use of volatile samples, such as OBM. A special cryogenic system can be installed that uses liquid

nitrogen to freeze volatile samples, but in the case of this work it introduced a source of vibrations

that made it impossible to achieve high magnification. Although, an Energy Dispersive X-Ray

analysis (EDX) was used to confirm deposition of calcium carbonate within a filter cake following

HPHT filtration of a 5 wt% carrier invert. Freshly prepared carrier was filtered through an HPHT

press, and a small piece of a filter cake was quenched in liquid nitrogen and placed in a FEI Quanta

250 SEM for analysis.

2.7 Field testing

A total of six full-scale field tests were carried out at several locations near Rocky Mountain

House, Alberta, Canada. The wells were drilled in pairs using the same double-type rig and mud

program. Despite passing through the same lithological sequences, the wells were separated into

four groups based on their surface location in order to further account for slight variations in the

geology, as shown in Figure 9. Table 2-3 provides an overview of the test and control wells.

Figure 9: Surface locations of the control and test well groups A, B, C, and D.

Page 48: A Field Application of Nanoparticles For Improved Downhole

36

Table 2-3: An overview of control and test wells.

Well Group LSD TD,

m MD Type

OBM

System

Initial Invert

Depth, m MD

Test A1 A 02-08-041-06-W5 3610 HZ** Cutter-D

90/10 613

Control A1 A 01-08-041-06-W5 3620 HZ Cutter-D

90/10 613

Control A2 A 02-10-041-06-W5 3250 HZ Cutter-D

90/10 613

Control A3* A 01-10-041-06-W5 3475 HZ Cutter-D

90/10 614

Test B1 B 02-21-041-07-W5 3516 HZ Cutter-D

85/15 613

Test B2 B 03-21-041-07-W5 3500 HZ Cutter-D

85/15 612

Control B1 B 12-19-041-07-W5 3705 HZ Cutter-D

85/15 613

Control B2 B 13-19-041-07-W5 3655 HZ Cutter-D

85/15 613

Test C1 C 04-35-037-08-W5 4033 HZ

Cutter-

D/D822

85/15

1201

Control C1 C 16-35-037-08-W5 3874 HZ

Cutter-

D/D822

85/15

1170

Control C2 C 09-36-037-08-W5 3905 HZ

Cutter-

D/D822

85/15

1205

Control C3 C 15-35-037-08-W5 3861 HZ

Cutter-

D/D822

85/15

1200

Control C4 C 03-35-037-08-W5 4097 HZ

Cutter-

D/D822

85/15

1198

Test D1 D 12-02-039-05-W5 3912 HZ

Cutter-

D/D822

85/15

612

Test D2 D 11-02-039-05-W5 3935 HZ

Cutter-

D/D822

85/15

613

* Includes a residual NP-based drilling fluid from well Test A1.

** HZ = Horizontal.

Page 49: A Field Application of Nanoparticles For Improved Downhole

37

2.7.1 Field test implementation

A custom carrier emulsion was mixed according to the procedure described above and delivered

to a rig site where it was transferred to an empty storage tank. Initially, 10 vol% of carrier emulsion

with respect to the total volume of circulating mud (typically 6–7 m3) were slowly added to the

active system over several circulations to achieve the target concentration of CNP of 0.5 wt%. As

drilling progressed, and more drilling fluid was transferred from a tank farm to maintain the

circulation volume, 1 m3 of carrier invert was used for every 10 m3 of pre-mix. A diagram of field

implementation is shown in Figure 10.

Figure 10: A schematic of NPs field implementation using a carrier approach.

20 m3 carrier

5 wt% CNP

9 m3 invert1 m3 carrier

10 vol% carrier

0.5 wt% CNP

Mixing Facility

Tanker Truck

Rig Tank Farm

Active Ri g Tanks Pre-mix Tank

Page 50: A Field Application of Nanoparticles For Improved Downhole

38

2.7.2 Field data collection

Field trials involved continuous presence at a rig site from initial depth to TD, and the data was

acquired throughout a day at specific depth increments (typically 200–400 m). At each depth point,

samples of drilling fluid were collected at a flow line and tested on-site. Mud weight, electrical

stability, retort analysis, and HPHT fluid loss were measured using the same lab techniques and

equipment as previously described. Subsequently, the necessary volume measurements and

calculations were performed to estimate mud losses. Additional observations included analysis of

cuttings and monitoring of drilling parameters to ensure that NPs did not cause any problems.

Ideally, extensive field analysis requires additional information such as wireline logs,

production data, and well testing data, which could be used to accurately determine the impact of

NPs on various aspects of drilling, completion or production. Unfortunately, due to the sensitive

and proprietary nature of such detailed information, an operator company could only provide

historic data from several offset wells in the area. The data was limited to mud losses and basic

properties of drilling fluids, and it was used to establish average values in control wells for

comparison.

Total mud losses were estimated using a conventional volumetric balance approach. The

method was based on calculating the difference between control volumes at two different depth

points. Given that no new volume was generated (e.g. product addition or new mud shipment), the

difference corresponded to the volume of total mud losses. However, this approach suffered from

a number of significant limitations and gave rise to large experimental errors, as will be explained

in further detail. A method for calculation of mud losses is shown in Figure 11.

Page 51: A Field Application of Nanoparticles For Improved Downhole

39

Figure 11: A diagram explaining calculation of mud losses in the field.

Control volume consisted of several discreet elements: a tank farm, a pre-mix tank, active

tanks, and hole volume. Consequently, each element needed to be determined separately. Tank

volumes were provided by floating sensors and therefore depended on their calibration and

accuracy. Volumes provided by tank sensors did not always correspond to the true volumes of

fluid delivered on site, hence why starting volumes used in the calculation of losses were based on

a sensor reading rather than the actual volume of fluid delivered. Figure 12 below shows one such

example of the discrepancy between tank volume readings provided by digital sensors as opposed

to a manual gauge.

Page 52: A Field Application of Nanoparticles For Improved Downhole

40

Figure 12: Example of the discrepancy between manual and digital floating sensors in a mud

storage tank.

Conversely, pipe and annular volumes can only be estimated theoretically. Calculations are

based on an inner diameter of individual hole sections, as well as on a capacity and displacement

of separate elements of a drilling string. While the information about drilling string assembly was

readily available from drilling reports, hole diameter could not be determined directly without a

calliper log. Therefore, it was assumed to be either equal to a bit size (e.g. no washout). It becomes

obvious than in the situations where hole washout is taking place, the actual hole volume may be

greater than the theoretical. This introduces errors in the calculation of losses, which often remain

unnoticed.

Another factor affecting the accuracy of mud losses calculation is associated with volume

increase due to products addition. For example, large amounts of barite could be added to increase

mud weight between two depth points of interest. This addition exaggerates the control volume,

Page 53: A Field Application of Nanoparticles For Improved Downhole

41

so that the difference in the total volumes between data points becomes smaller and no longer

reflects losses that occurred. While it is possible to correct for such volume increase, the mud

company managing drilling fluids at the test locations did not adopt that practice. Therefore, the

effect of product additions was ignored in order to maintain consistency with the historic data.

While the volumetric balance approach employed in this thesis provided an estimate of

whole mud losses, it did not distinguish between surface and downhole losses. Surface losses were

mostly caused by residual mud on discarded drilled cuttings and centrifuge solids. As a result, they

varied depending on the efficiency of shakers, mesh size, ROP, formation drilled, type of a drill

bit, amount of operating hours and settings of a centrifuge, etc. Since only some of this information

was available, it was impossible to accurately determine the relative ratio of surface and downhole

losses for each well. Therefore, it was assumed that each well on average experienced similar

geology, mud system, drilling parameters, and solids control protocol throughout its life. The

second assumption was that NPs only affected the downhole portion of total mud losses. As a

result, any change in the final mud losses averaged across control and test wells could be attributed

to the action of NPs.

Page 54: A Field Application of Nanoparticles For Improved Downhole

42

CHAPTER THREE: RESULTS AND DISCUSSION

3.1 Carrier emulsion

3.1.1 Properties of a carrier emulsion

The carrier emulsion approach of introducing NPs to a drilling fluid was based on volumetric

dilution. Consequently, it was desirable to achieve high concentration of NPs in a carrier fluid,

which in turn would reduce the volume required to achieve a target concentration of NPs in a host

fluid. However, even when aqueous precursors are near-saturated, the maximum concentration of

NPs is limited by the amount of water that the invert emulsion can support before breaking down.

Laboratory observation suggested that carrier emulsion separated at concentrations of CNP above

5 wt%. Accordingly, since previous lab experiments set the optimum concentration of NPs in a

host fluid at 0.5 wt%, such approach involved a 10-fold dilution. As mentioned earlier, it was

important to minimize impact of NPs addition on basic properties of a drilling fluid. Therefore,

formulation of carrier emulsion was varied until its final properties approximated those of a typical,

most representative diesel-based mud. Laboratory experiments suggested that the impact of

addition of carrier emulsion on basic properties of several different commercial mud samples was

insignificant. This rendered the carrier emulsion approach a valid alternative to the direct, in situ

method, as will be explored in further detail in the following sections.

The final lab formulation of a carrier emulsion with 5 wt% CNP was recreated on a 20 m3

volume scale according to the procedure described in Chapter Two. Sample of a product was

collected at the end of the mixing process, and its properties were evaluated against the lab-based

standard. Properties of the final formulation of carrier invert samples prepared on a lab and field

scale are provided in Table 3-1. The results indicate that the large-scale process did not affect the

apparent properties of a carrier emulsion.

Page 55: A Field Application of Nanoparticles For Improved Downhole

43

Table 3-1: Comparison of the properties of the carrier emulsion samples with 5 wt% CNP

prepared on a lab and field scale.

Property Lab

Scale

Field

Scale

MW (kg/m3) 1049±3 1045±5

ES (V) 780±20 582±57

PV (cP) 12±2 12±2

YP (lb/100 ft2) 5±2 5±2

GS 10 s (lb/100 ft2) 4±1 4±1

GS 10 min (lb/100 ft2) 6±1 5±1

OWR 83/17 83/17

Solids (vol%) 8±1 8±1

To further examine the applicability of the carrier emulsion approach, it was important to

evaluate long-term stability of the final formulation of a carrier emulsion. During normal drilling,

mud in storage tanks can remain stagnant up to several days. This leads to gravity segregation and

consolidation of solids at the bottom of a tank. Typically, when additional drilling fluid from the

storage tanks is required to make up volume, it is circulated for several hours using built-in pump

to homogenize the mixture. Therefore, it was essential to monitor long-term stability of carrier

emulsion samples and ensure that they can perform in field conditions. Samples of carrier invert

prepared on a lab and field scale were stored at ambient conditions for four days in order to

determine their stability, as shown in Figure 13. Observations suggest that both carrier invert

preparations exhibited similar rate of settling as a typical diesel-based invert emulsion drilling fluid

containing conventional LCM and other solids. After 96 hours of storage under static, ambient

Page 56: A Field Application of Nanoparticles For Improved Downhole

44

conditions, all three samples showed a clear solid phase boundary. However, even following one

week of settling, solids could be easily dispersed by shaking, yielding a once again homogeneous

emulsion in all samples. In addition, despite having similar composition, the colour of carrier

emulsion samples was lighter than that of a control drilling fluid. This indicates the presence of

white calcium carbonate precipitate and confirms that the coprecipitation reaction took place.

While this section focused on the macroscopic properties of lab- and field-scale carrier

emulsion samples, it was also necessary to investigate the process on a microscopic level. This is

the topic of the next two sections.

Figure 13: Long-term stability of carrier emulsion samples with 5 wt% CNP prepared in a

laboratory vs field scale. A typical commercial control invert is shown for comparison.

3.1.2 DLS analysis of a carrier emulsion

Dynamic Light Scattering is one of the most commonly used analytical techniques for NPs

characterization (Murty et al., 2013). Analysis is usually based on five different parameters that

are used together and provide an estimate of the average particle size in a sample. The calculation

Page 57: A Field Application of Nanoparticles For Improved Downhole

45

of the parameters involves complex cumulants and distribution fitting algorithms, which are

sensitive to the concentration and polydispersity of a sample. While the actual theory involved in

DLS is outside the scope of this thesis, it is important to stress that in an ideal, monodispersed

system, intensity-, volume-, number-, and Z-average sizes should fall within a close range, and

polydispersity index (PDI) should be low. Conversely, when large, sedimenting or agglomerating

particles are present in a sample, data quality deteriorates. Usually, this leads to a large deviation

between the calculated parameters and results in a high PDI. For that reason, DLS analysis of

invert emulsion drilling fluids is problematic due to a wide range of particle sizes, in addition to

the presence of water pools and micelles.

Results of the DLS analysis on the lab- and field-based carrier emulsion samples are

provided in Table 3-2. One the one hand, the data shows some disagreement between the calculated

parameters. For instance, the mean particle size in the lab sample ranged between 150±80 nm and

450±80 nm, depending on the method used. While a number of factors could contribute to this

deviation, sample polydispersity was likely the main cause. On the other hand, standard deviation

between separate replicates falls within 10–20%, which is acceptable, considering a rather broad

size distribution (PDI 0.3–0.4). Furthermore, the data for the two samples was in a good agreement,

which also suggests that the scale-up process did not significantly affect the average particle size.

Table 3-2: Comparison of DLS data on carrier emulsion samples with 5 wt% CNP prepared in a

laboratory vs field scale. 25 °C, THF dispersant.

Carrier

Scale

Intensity

Mean (nm)

Volume

Mean (nm)

Number

Mean (nm)

Z-Average

(nm) PDI

Lab 280±20 300±40 270±20 450±80 0.400±0.050

Field 300±60 280±30 150±80 395±30 0.360±0.060

Page 58: A Field Application of Nanoparticles For Improved Downhole

46

To further investigate the effect of carrier preparation on the particle size, intensity

distributions of the two samples were plotted for comparison. As shown in Figure 15 below, the

lab sample exhibited a range of sizes between 150–400 nm, as opposed to 50–900 nm of the field

sample. While this may be caused by an experimental error, it may also indicate that high shear

mixing of the lab sample provided better size control than the high-pressure shearing employed in

the preparation of the field sample. Considering that both samples had the exact same components,

their solid fraction was primarily composed of CNP. As a result, changes in the DLS data were

likely to reflect the difference between two preparation methods.

Despite the aforementioned limitations of the DLS analysis of carrier emulsions, it

provided a helpful insight when combined with other experimental data.

Figure 14: Intensity size distribution in a carrier emulsion samples with 5 wt% CNP produced in

a laboratory (A) vs field (B).

3.1.3 SEM and EDX analysis of a carrier emulsion

Scanning Electron Microscopy is an ideal method to investigate size, shape, and distribution of

sub-micron particles (Murty et al., 2013). However, most typical applications involve dry powders,

which excludes invert emulsion drilling fluids from the list of suitable samples. Instead, SEM was

used to gain insight into the distribution of solids within a filter cake formed during static HPHT

Page 59: A Field Application of Nanoparticles For Improved Downhole

47

fluid loss experiment. Since the cake was oil-wet, cryogenic sample probe and chamber were

employed to limit volatility of the sample. Unfortunately, at magnification levels above 10,000x,

vibrations from the cryogenic unit made it impossible to maintain steady focus, and the image

became blurry.

To further facilitate the analysis, Energy Dispersive X-ray (EDX) spectroscopy was

employed to investigate elemental composition of the sample. EDX distinguishes between

elements based on their unique atomic structure and can be tuned to detect elements of interest. In

this case, the elements included those added as part of the precursors: calcium, carbon, oxygen,

potassium, and chlorine. In addition to indicating relative amounts of different elements within an

area of a sample, EDX is also capable of mapping that data as separate colored images for each of

the element. As a result, such images can indicate how different elements are correlated via

chemical bonding, as well as how they are distributed within a focus area. In order to make such

links more obvious, EDX mapping images were overlaid using different levels of transparency.

Black background corresponds to the lack of the selected elements, while the colored foreground

becomes more intense with increasing concentration.

An SEM image of the profile of a filter cake at 50x magnification is provided in Figure 15-

A, while the corresponding EDX mapping is shown in Figure 15-B below. It is worth pointing out

that the temperature of the sample was maintained at -150 C, which caused cracks and may have

otherwise affected the structure of the cake. The analysis of EDX images suggested that all five

elements of interest were distributed uniformly throughout the cake. However, higher levels of

magnification were required to discern any structural details.

Page 60: A Field Application of Nanoparticles For Improved Downhole

48

Figure 15: SEM image of the profile of the filter cake formed by a lab sample of carrier

emulsion with 5 wt% CNP (A); the corresponding overlapped EDX images showing uniform

distribution of calcium, carbon, oxygen, potassium, and chloride (B).

Two different areas within the sample were chosen to attempt further magnification, as

shown on the left- and right-hand side of Figure 16, respectively. The first area was acquired at

5,000x magnification and showed several large irregular-shaped particles and a number of

scattered rhombic crystals. The corresponding EDX mapping (shown in Figure 16-C) revealed that

the crystals were in fact potassium chloride, which was a by-product of the co-precipitation

reaction. Similarly, overlaid images in Figure 16-B also showed that calcium, carbon, and oxygen

were closely correlated. While this confirms the formation of calcium carbonate during the co-

precipitation reaction, it does not provide enough resolution to observe NPs. The large structural

Page 61: A Field Application of Nanoparticles For Improved Downhole

49

features featured in the SEM image are either entirely composed of calcium carbonate or, which

is more likely considering the atypical shape, present clay platelets or agglomerates coated with

CNP. Colored spots in the background of the EDX image correspond to the parts of the SEM

image where no individual particles are visible at this level of magnification.

Figure 16: SEM images of two different areas within the profile of a filter cake formed by a lab

sample of carrier emulsion with 5 wt% CNP (A, D); the corresponding overlapped EDX images

showing correlation between calcium, carbon, and oxygen (B, E); the same areas showing

correlation between potassium, and chlorine (C, F).

Subsequently, the second area within the sample was chosen and examined at higher

magnification of 10,000x, as shown in Figure 16-D. Similar to the image on the left, SEM showed

a number of crystals ranging between 1–3 m, but the resolution was still too low to notice any

Page 62: A Field Application of Nanoparticles For Improved Downhole

50

details in the background. EDX analysis shown in Figure 16-F verified that crystals were formed

by potassium carbonate, likely as a result of high temperature and pressure conditions during the

HPHT test. Accordingly, mapping images in Figure 16-E indicated the presence of calcium

carbonate and showed dark spots corresponding to crystals of potassium chloride. However, the

background of the image suggested the presence of sub-micron particles that constituted the cake

matrix but were too small to observe on that scale.

All further attempts to increase magnification and resolution were unsuccessful due to the

loss of focus; therefore, the experiment only provided limited information. The only conclusion

was that the emulsion-based chemical co-precipitation did in fact produce calcium carbonate, and

that such particles were primarily in the sub-micron domain.

3.2 Basic properties of commercial drilling fluids with CNP

Before probing the potential benefit of NPs on filtration of drilling fluids, it was first necessary to

ensure that their addition did not affect basic properties of mud to a significant extent. Both the in

situ and the carrier emulsion methods were investigated, and the results suggested that neither

exhibited adverse effects on the samples. Properties of commercial virgin and recycled drilling

fluids in the presence of 0.5 wt% CNP were evaluated following the procedures outlined in Chapter

Two.

Theoretically, mud weight of mud samples should increase by less than 1% in the case if

in situ method, while the corresponding increase in the case of emulsion approach depends on how

closely densities of the two fluids are matched. In reality, however, the difference in mud weight

measured with a mud balance was so negligible, that for the most part it was obscured by the

Page 63: A Field Application of Nanoparticles For Improved Downhole

51

experimental error. This is shown in Figure 17 for virgin and recycled mud samples (left and right,

respectively).

Figure 17: Impact of 0.5 wt% CNP on mud weight of commercial invert emulsion drilling fluids

using in situ and carrier emulsion methods.

The effect of addition of 0.5 wt% CNP on electrical stability of commercial drilling fluids

is shown in Figure 18. The in situ method either did not affect ES of the samples, or its impact was

minimal. This result was expected considering that composition of the host fluid was altered only

marginally. On the other hand, carrier emulsion method reduced electrical stability of drilling

fluids slightly. The effect was most pronounced in the case of virgin mineral and diesel OBM,

while the other samples were not affected to the same extent. This behaviour has not yet been fully

understood. However, electrical stability was in the acceptable range in all samples, and it was

deemed that NPs did not result in any dramatic differences

Page 64: A Field Application of Nanoparticles For Improved Downhole

52

Figure 18: Impact of 0.5 wt% CNP on electrical stability of commercial invert emulsion drilling

fluids using in situ and carrier emulsion methods.

Plastic viscosity and yield point of drilling fluid samples are provided in the top and bottom

of Figure 19, respectively. Both parameters were affected by the addition of 0.5 wt% NPs, and

similar to the previous discussion, carrier emulsion approach exhibited more significant effect.

The general trend suggested increase in PV and YP in the presence of CNP. However, despite the

differences in rheological properties, they were well within the operating range maintained during

drilling operation.

Page 65: A Field Application of Nanoparticles For Improved Downhole

53

Figure 19: Impact of 0.5 wt% CNP on plastic viscosity (A, B) and yield point (C, D) of

commercial invert emulsion drilling fluids.

To further determine the potential impact of NPs on rheology of mud samples, gel strengths

at 10 s and 10 min were measured and summarized in Figure 20. Similar to the case of plastic

viscosity and yield point, impact on gel strengths in most drilling fluids was minimal.

Page 66: A Field Application of Nanoparticles For Improved Downhole

54

Figure 20: Impact of 0.5 wt% CNP on gel strength of commercial invert emulsion drilling fluids

at 10 s (A, B) and 10 min (C, D).

Last but not least, retort analysis was used to measure volumetric composition of drilling

fluids with and without NPs. The results shown in Figure 21 suggest that the impact on OWR and

solids content of all mud samples was less than 1%, which again was within the normal operating

range.

Once the experiments confirmed that NPs at 0.5 wt% can be safely introduced to typical

commercial drilling fluids without causing significant changes in the final properties, it was

appropriate to focus laboratory testing on fluid loss.

Page 67: A Field Application of Nanoparticles For Improved Downhole

55

Figure 21: Impact of 0.5 wt% CNP on volumetric composition of virgin (A) and recycled (B)

commercial invert emulsion drilling fluids.

3.3 Fluid loss in commercial drilling fluids with CNP

Zakaria et al. (2012) showed that in situ calcium carbonate NPs at 0.5 wt% were effective reducing

spurt loss and filtrate volume in several types of commercial invert emulsion drilling fluids. The

experiments were independently reproduced in the lab using a modified in situ preparation method

based on aqueous precursors. Three different types of virgin drilling fluids were tested, as

presented in the left-hand side of Figure 22.

Page 68: A Field Application of Nanoparticles For Improved Downhole

56

Figure 22: Impact of 0.5 wt% CNP on HPHT fluid loss (A, B) and filter cake thickness (C, D)

of commercial invert emulsion drilling fluids at 80 °C and 500 psi.

Both virgin diesel-based fluids showed reduction of fluid loss in the presence of 0.5 wt%

CNP. It is worth mentioning that virgin drilling fluids typically exhibit much higher fluid loss than

their conditioned, field-based counterparts. As a result, this makes any changes in the data more

apparent. Subsequently, virgin samples of Cutter-D and Diesel OBM showed a 17% and 33%

reduction of average filtrate volumes at 30 min in the presence of in situ 0.5 wt% CNP,

respectively. The corresponding average filter cake thickness was also reduced compared to the

control samples.

On the one hand, the HPHT values in the virgin Cutter-D sample were further reduced by

52% when the same concentration of CNP was introduced via a carrier emulsion approach. On the

Page 69: A Field Application of Nanoparticles For Improved Downhole

57

other hand, the in situ and carrier emulsion methods exhibited similar performance in the case of

the virgin diesel-based OBM. The difference was attributed to the presence of clays alongside NPs

in the carrier emulsion approach. It is possible that NPs interacted with clay platelets and formed

a composite network that reinforced the structure of the cake and reduced its permeability. An

example of such process was investigated and reported elsewhere (Walz, 2011).

Finally, regardless of the approach utilized, CNP did not affect the filtration properties of

the virgin mineral-based OBM, where HPHT values and cake thickness remained within the

standard deviation of the control sample. This suggested that adverse interactions likely took place

between CNP and other components in the mud, which in turn limited the ability of NPs to reduce

filtration rates.

Since virgin drilling fluids did not fully represent properly conditioned and matured field

mud, it was essential to further investigate the impact of a composition of recycled samples on its

filtration properties in the presence of NPs. The average values of filtrate volume and cake

thickness are provided on the right-hand side of Figure 22. Field samples typically contain higher

concentration of filtration control agents, as well as higher solids content in general. Accordingly,

the HPHT values in all three recycled control samples were in the range of 4–6 mL2, which was

much lower than in the case of the virgin fluids. However, despite the already optimized

composition of each of the three field samples, CNP at 0.5 wt% reduced the filtrate volumes even

further. Unlike the virgin fluids, in situ and carrier emulsion methods exhibited negligible

difference and achieved an average of 30–40% lower filtrate volumes compared to the control

samples. In addition, the average thickness of a filter cake in field fluids was also reduced in the

presence of 0.5 wt% CNP, which is a significant result.

Page 70: A Field Application of Nanoparticles For Improved Downhole

58

To summarize, the results of the HPHT experiments demonstrated that CNP at 0.5 wt%

improved filtration control properties of virgin and real-life invert emulsion drilling fluids. While

the performance of NPs depended on the nature of the host drilling fluids as well as the method of

CNP synthesis, average HPHT values and the corresponding cake thickness were reduced in the

most representative field samples. The reduction was consistent with the filtration model

developed in Chapter One. More importantly, it was also shown that the carrier emulsion approach

achieved the same reduction in recycled fluids as the in situ approach. This signified that the

process could be scaled up to a field application, which is the topic of the remainder of this chapter.

3.4 Field testing

Before proceeding with the analysis of field data, it is important to discuss its accuracy and

limitations. As was pointed out earlier, typical field data suffers from large degree of errors that

are caused both by the limitations of the measurement devices as well as the human factor. In

addition, total mud losses are normally affected by so many different parameters, that they can

vary dramatically even when wells are drilled in the same location and using the same mud system.

Since the objective of the field trial was to determine the impact of NPs on mud losses, a reliable

baseline for comparison had to be established first. Ideally, large volumes of field data for control

and test wells are required to reduce statistical error and provide a more accurate estimate.

However, in the case of this thesis the amount of available information was rather limited. In order

to increase reliability of the field results, the data was collected independently and then compared

with measured and calculated values provided by a mud company representative. In all cases the

difference was within ±1%, so the results are used interchangeably throughout this chapter.

Finally, despite the use of proper scientific methodology and a consistent approach, it must be

Page 71: A Field Application of Nanoparticles For Improved Downhole

59

acknowledged that the field data presented herein serves to provide general and relative trends

rather than absolute values.

3.4.1 HPHT fluid loss

Results of HPHT fluid loss experiments in all control and test wells are shown in Figure 23. Once

WBM is displaced, invert emulsion drilling fluids require time to fully develop their properties as

they are continuously circulated, sheared, and heated. Furthermore, when a fresh drilling fluid is

used, it first must be conditioned with additives, as specified in a mud program. Additives are

typically added slowly, over several circulations; hence their impact on HPHT is not immediately

apparent. This effect is often observed as a trend towards lower HPHT values with increasing

depth, until a characteristic plateau region is reached.

Page 72: A Field Application of Nanoparticles For Improved Downhole

60

Figure 23: HPHT fluid loss as a function of measured depth in the control and test wells at 80

°C and 500 psi.

Since the well group A contains control, depleted, and test wells, it allows to investigate

the effect of concentration of NPs on HPHT values. As shown in Figure 23-A, Control A1 started

at 6.5 mL×2 filtrate volume, which then decreased to 4.8 mL×2 at approximately 1500 m MD and

remained constant until TD. Conversely, Test A1 started at a lower value of 4 mL×2, which then

stabilized at 1 mL×2. This supports the notion that the concentration of NPs requires time to reach

the optimum values, especially when fresh invert is used. Similarly, Control A3*, which already

contained NPs at a working concentration, showed even lower initial HPHT fluid loss of 3 mL×2.

After sufficient circulation, it reached the same characteristic value of 1 mL×2 as in the case of

Page 73: A Field Application of Nanoparticles For Improved Downhole

61

Test A1. As drilling progressed, drilling fluid was diluted with fresh invert, and the concentration

of NPs was depleted. As a result, HPHT values increased to 2 mL×2 near the end of the well.

While the control HPHT data for group B was not available, wells Test B1, Test B2 showed

low fluid loss of 1.5 mL×2, which is consistent with the well Test A1.

The four control wells in group C showed a wide range of HPHT values that varied between

2–8 mL×2, while the HPHT values in well Test C1 exhibited a trend similar to that in Test A1 and

reached low values of 2–3 mL×2 in the presence of NPs. This also holds for both test wells in

group D, where fluid loss remained between 1–3 mL×2.

The above HPHT data can be summarized and represented as the average fluid loss within

each well group, which is shown in Figure 24. Results clearly indicate that on average, NPs

reduced HPHT filtrate volumes by 20–30%, compared to control wells that contained conventional

LCM alone. This correlates well with the lab experiments discussed in the previous section, where

the average reduction of filtrate volume of 30% was observed in a similar type of drilling fluids.

Figure 24: Average HPHT fluid loss in the control and test wells at 80 °C and 500 psi.

Page 74: A Field Application of Nanoparticles For Improved Downhole

62

3.4.2 The effect of concentration of carrier

The effect of the concentration of carrier emulsion on HPHT values in the test wells was

investigated further. The theoretical concentration of carrier at each depth point was calculated

using a volumetric ratio of carrier to the total circulating mud volume and then plotted against

measured depth, as shown in Figure 25. Recall that the concentration of NPs in carrier was 5 wt%,

which corresponds to a 10 vol% concentration in the circulating mud system. Wells Test A1 and

Test C1 show that as the volumetric ratio of carrier increased towards 10% v/v, HPHT values

decreased accordingly. Once the concentration of NPs in the drilling fluid was maintained within

the target range, HPHT filtrate volumes remained constant. Similar conclusions can be drawn for

groups B and D, which started at higher concentrations of carrier. As a result, the change in HPHT

values was not as dramatic as in the case of groups A and C. This demonstrates that the

concentration of NPs in a drilling fluid had a direct impact on its filtration control properties.

Page 75: A Field Application of Nanoparticles For Improved Downhole

63

Figure 25: HPHT fluid loss and calculated concentration of carrier as a function of measured

depth in the test wells.

3.4.3 Cumulative mud losses

Cumulative mud losses while drilling in control and test wells are presented as a function of

measured depth in Figure 26. Comparison between test and control wells suggests that NPs had an

impact on mud losses. As such, the well Test A1 showed the lowest total mud losses within the

group, with less than 60 m3 lost at TD. Control A3*, which demonstrated the effect of leftover NPs

in HPHT experiments, also showed lower losses at TD compared to control wells that only

contained conventional LCM. Similar situation can be observed in group B, where both test wells

suffered lower mud losses while drilling than the control wells. Group C demonstrates a wide

range of losses in control wells, which vary between 70 and 117 m3 at TD. However, Test C1 with

Page 76: A Field Application of Nanoparticles For Improved Downhole

64

65 m3 at TD still falls on the lower side that range. Unfortunately, no control data for group D was

available, therefore the direct comparison cannot be made within the same group.

Figure 26: Cumulative mud losses as a function of measured depth in the control and test wells.

To summarize the data shown in Figure 26 above, final cumulative mud losses while

drilling are provided in Figure 27. The results indicate that the losses in the presence of NPs were

on average 20–30% lower than in control wells. This range is very consistent with observations

made during HPHT fluid loss experiments. However, it should be noted that the final cumulative

losses were measured at TD, which varied slightly between the wells. In order to standardize the

data, cumulative losses were converted to losses per 100 m drilled, which are discussed in further

detail in the following section.

Page 77: A Field Application of Nanoparticles For Improved Downhole

65

Figure 27: Final cumulative mud losses at TD in the control and test wells.

3.4.4 Mud losses per 100 m drilled

Similar to the cumulative mud losses while drilling, losses per 100 m drilled showed lower values

in the presence of NPs. As can be seen in Figure 28, test wells consistently achieved on average

of 2 m3/100 m, which is lower than the control wells containing conventional LCM alone. In

addition, the test wells demonstrated more narrow variation of losses with depth, which further

supports the notion that NPs help to control downhole losses.

The data can also be represented as the average losses per 100 m drilled at TD, as shown

in Figure 29. The results clearly indicate that, even after correcting for variation in TD, test wells

still provided a 20–30% reduction of losses compared to the control wells.

Page 78: A Field Application of Nanoparticles For Improved Downhole

66

Figure 28: Mud losses per 100 m drilled as a function of measured depth in the control and test

wells.

Figure 29: Average mud losses per 100 m drilled at TD in the control and test wells.

Page 79: A Field Application of Nanoparticles For Improved Downhole

67

3.4.5 Correlation of HPHT fluid loss and mud losses

As mentioned previously, correlation of HPHT fluid loss and the actual mud losses remains one

of the challenges associated with field testing. Figure 30 shows a plot of HPHT filtrate volumes

versus the corresponding instantaneous mud losses that occurred between two successive depth

points. While the results do not form a consistent trend, test wells (filled markers) tend to be located

closer to the origin point of the graph than the control wells (open markers). In turn, this likely

indicates that reduction of HPHT values corresponds to lower mud losses in the presence of NPs.

Figure 30: Correlation between HPHT fluid loss and instantaneous mud losses in the test wells.

Page 80: A Field Application of Nanoparticles For Improved Downhole

68

CHAPTER FOUR: CONCLUSIONS, CONTRIBUTIONS AND RECOMMENDATIONS

4.1 Conclusions

The results of the lab and field experiments provided in this thesis can be summarized in the

following key points:

1) CNP were successfully synthesized in the lab using chemical co-precipitation of aqueous

precursors. A carrier alternative to the in situ preparation onsite was proposed to ensure

scale-up for a field application. A custom invert carrier emulsion with 5 wt% CNP was

synthesized using a reverse micelle approach and was then used to introduce NPs to a host

fluid via volumetric dilution.

2) Six different invert emulsion drilling fluids were tested in the presence of 0.5 wt% CNP

introduced via two different routes. The in situ method did not significantly impact the

basic mud properties, while the effect was more pronounced when carrier emulsion was

used at a 10 vol% ratio.

3) CNP helped to reduce average HPHT filtrate volumes in virgin and recycled samples by

30–40%. Carrier emulsion approach demonstrated similar performance as the in situ

method, which suggested that scale-up to a field application was possible.

4) Carrier emulsion approach was successfully implemented at an industrial drilling fluids

processing facility. The method was fully compatible with standard oilfield chemicals and

did not require significant deviation from routine mixing operations. A total of 6 batches

of carrier emulsion were successfully prepared on a scale of 20 m3 volume.

5) Macroscopic properties of carrier produced on a large scale were in good agreement with

the lab-based benchmark, which suggested that the scale-up process proceeded according

to the expectations. Lab- and field-scale carrier emulsion samples with 5 wt% CNP were

Page 81: A Field Application of Nanoparticles For Improved Downhole

69

stable after long-term storage. After one week of static ageing under ambient conditions,

samples showed the same extent of solids settling as a representative control sample of a

typical, commercial drilling fluid.

6) SEM and EDX analysis confirmed the distribution of calcium carbonate and potassium

chloride within the profile of a filter cake formed by a carrier emulsion. While the

resolution was insufficient to distinguish sub-micron particles, the experiments provided a

valuable insight into elemental composition of the sample.

7) DLS experiments on lab- and field-based carrier emulsion showed the average particle size

of approximately 300–400 nm. Intensity distribution data further suggested that the lab

preparation achieved a more narrow size range as opposed to the field sample, where the

particles covered the range of 50–900 nm. Since laboratory method employed high shear

mixing conditions, it may have provided better size control than the jet shear principle.

However, this is significant result that demonstrates that the reverse micelle synthesis of

NPs can be successfully implemented on a large scale. Furthermore, the process utilized

common oilfield chemicals and additives that are native to most drilling fluids. Therefore,

the process was economical and did not present significant health risks.

8) Six different commercial drilling fluids were tested in the lab to determine impact of CNP

on their properties. Virgin as well as recycled samples showed relatively insignificant

change in basic experimental parameters. As expected, the impact of in situ method was

less pronounced than the 10 vol% dilution required in the carrier emulsion approach.

However, all three samples of most typical, diesel-based OBM showed good compatibility

with the carrier formulation.

Page 82: A Field Application of Nanoparticles For Improved Downhole

70

9) Results of the HPHT fluid loss experiments on commercial invert emulsion drilling fluids

in the presence of in situ 0.5 wt% CNP were consistent with the previous work by Zakaria

et al. (2012). Filtrate volumes and the corresponding filter cake thickness were reduced by

20–40% on average. Subsequently, the experiments were recreated using the carrier

emulsion approach instead. In the case of the recycled samples, both methods achieved the

same reduction of HPHT values on the order of 30–40%.

10) Six field tests were conducted in the province of Alberta, Canada in 2014. A carrier

emulsion approach was used in all wells, which were assigned into four groups to eliminate

effect of their surface location. Field trials involved regular mud testing, calculation of mud

losses, analysis of cuttings, and continuous monitoring of drilling operation.

Simultaneously, a mud company representative acquired the same data to ensure its

reliability. Comparison revealed that the differences were within ±1%.

11) The basic properties of drilling fluids in the test wells were within same range as in control

wells that contained conventional LCM, while the HPHT values were on average 20–30%

lower. The results were consistent with the lab experiments performed on a small-scale

carrier emulsion. Furthermore, correlation between filtrate volumes and the calculated

concentration of NPs in the whole mud was established.

12) A volumetric balance approach was implemented to evaluate total mud losses in the test

wells. The data was compared to the mud losses in historic offset wells provided by a mud

operator. Despite the limited sample size and high degree of errors, test wells consistently

showed 20–30% lower cumulative losses at TD. Accordingly, losses per 100 m were also

lower than in control wells.

Page 83: A Field Application of Nanoparticles For Improved Downhole

71

13) The results of lab and field experiments were consistent with the model of filtration in the

presence of NPs developed in Chapter One. Lower filtrate volumes corresponded to a

thinner filter cake, which could only mean that CNP were effective reducing the

permeability of the medium.

4.2 Contributions to knowledge

The work presented in this thesis produced significant results that will be of interest to a wide

audience, including the drilling industry and the nanotechnology researchers alike. Firstly it was

demonstrated that a microemulsion-based, in situ synthesis of NPs can be successfully

implemented on an industrial scale. A carrier emulsion of 20 m3 volume containing 1,050 kg of

CNP was produced in a specialized mixing facility following a bench-scale technique and

exhibited similar average particle size and final properties as the samples prepared in the

laboratory. Secondly the results suggested that the carrier emulsion approach can be incorporated

into routine drilling operations without causing any problems, delays or health concerns.

Subsequently, analysis of mud losses indicated that the test wells containing 0.5 wt% CNP

achieved on average 20–30% lower cumulative losses while drilling compared to the control wells

using typical invert emulsion drilling fluids. This bridged the gap between laboratory-based HPHT

experiments and the actual mud losses and confirmed the performance of CNP under real-life

conditions.

4.3 Recommendations for future work

In conclusion, a novel approach for a large-scale manufacturing of NPs was successfully

developed and extensively tested with a variety of commercial drilling fluids. Lab and field

Page 84: A Field Application of Nanoparticles For Improved Downhole

72

experiments provided promising preliminary results and demonstrated potential of the carrier

emulsion approach for reducing downhole mud losses. This opens a wide range of possible

directions for future research efforts. Several most important recommendations are listed below:

1) Modify the carrier emulsion approach to increase the maximum concentration of CNP in

order to reduce the dilution ratio and make the process more cost-effective.

2) Investigate the possibility of synthesis of other types of NPs using the carrier emulsion

approach and perform extensive lab testing to determine their impact on essential mud

properties.

3) Conduct additional field trials to increase the sample size and reduce the statistical error in

the data. If possible, request wireline logs, tower sheets, archive of Pason Autodriller data,

and other helpful information that would allow to expand the focus of the data analysis

beyond simple monitoring of total mud losses.

Page 85: A Field Application of Nanoparticles For Improved Downhole

73

References

Agarwal, S., Walker, M. L., Prieve, C. D. & Soong, Y., “Using Nanoparticles and Nanofluids to

Tailor Transport Properties of Drilling Fluids for HPHT Operations.” AADE NTCE-18-05,

(2009).

Al-Hitti, H. A., Al-Assaf, A. S. & Abrahim, S. D., “Reduction of Formation Damage Due to

Drilling Muds.” Journal of Engineering, 11(1), 21–32, (2005).

Al-Riyama, K. & Sharma, M. M., “Filtration Properties of Oil-in-Water Emulsion Containing

Solids.” SPE Drilling & Completion, 164–172, (2004).

Amanullah, M., Al-Arfaj, M. & Al-Abdullatif, Z., “Preliminary Test Results of Nano-based

Drilling Fluids for Oil and Gas Field Application.” Presented at SPE/IADC Drilling

Conference and Exhibition, Amsterdam, Netherlands, (2011).

Amanullah, M. & Al-Tahini, A., “Nano-Technology- Its Significance in Smart Fluid

Development for Oil and Gas Field Application.” Presented at SPE Saudi Arabia Section

Technical Symposium and Exhibition, Alkhobar, Saudi Arabia, (2009).

API RP-13B, “Recommended Practice for Field Testing Oil-Based Drilling Fluids”, pp. 1–141,

(2014).

ASME Shale Shaker Committee, “Drilling Fluids Processing Handbook”, pp. 257–282, (2005).

http://doi.org/10.1016/B978-075067775-2/50012-3

Bennion, B. D., Thomas, B. F., Jamaluddin, A. K. M., Ma, T. & Agnew, C., “Formation Damage

and Reservoir Considerations for Overbalanced and Underbalanced CT Operations.”

Presented at 6th International Conference on Coiled Tubing Technologies, Houston, USA,

(1997).

Bezeme, C. & Havenaar, I., “Filtration Behavior of Circulating Drilling Fluids.” SPE Journal

Trans., 6(4), 292–298, (1966).

Boyd, J., Parkinson, C. & Sherman, P., “Factors Affecting Emulsion Stability and the HLB

Concept.” Journal of Colloid and Interface Science, 41(2), 359–370, (1972).

Caenn, R., Darley, H. C. H. & Gray, G. R., “Composition and Properties of Drilling and

Completion Fluids”, pp. 1–720, Gulf Publishing Company, (2011).

http://doi.org/10.1016/B978-0-12-383858-2.00001-9

Cai, J., Chenevert, M. E., Sharma, M. M. & Friedheim, J., “Decreasing Water Invasion Into

Atoka Shale Using Nonmodified Silica Nanoparticles”, SPE Drilling & Completion, 27(1),

103–112, (2012).

Page 86: A Field Application of Nanoparticles For Improved Downhole

74

Cao, G., “Nanostructures & Nanomaterials: Synthesis, Properties and Applications”, Imperial

College Press, London, UK, (2004).

Cargnel, D. R. & Luzardo, P. L., “Particle Size Distribution of CaCO3 in Drill-in Fluids: Theory

and Applications.” Presented at SPE Latin American and Caribbean Petroleum Engineering

Conference, Caracas, Venezuela, (1999).

Chelton, H. M., “Darcy’s Law Applied to Drilling Fluid Filtration”, SPE 1649, (1967).

Chin, W. C., “Formation Invasion With Applications to Measurement While Drilling, Time-

lapse Analysis, and Formation Damage”, Gulf Publishing Company, Houston, Texas,

(1995).

Contreras, O., Hareland, G., Husein, M., Nygaard, R. & Alsaba, M., “Application of In-House

Prepared Nanoparticles as Filtration Control Additive to Reduce Formation Damage.”

Presented at SPE International Symposium and Exhibition on Formation Damage Control,

Lafayette, USA, (2014a). http://doi.org/10.2118/168116-MS

Contreras, O., Hareland, G., Husein, M., Nygaard, R. & Alsaba, M., “Experimental Investigation

on Wellbore Strengthening in Shales by Means of Nanoparticle-Based Drilling Fluids.”

Presented at SPE Annual Technical Conference and Exhibition, Amsterdam, Netherlands,

(2014b).

Contreras, O., Hareland, G., Husein, M., Nygaard, R. & Alsaba, M., “Wellbore Strengthening in

Sandstones by Means of Nanoparticle-Based NPs Applications in Drilling Industry.”

Presented at SPE Deepwater Drilling and Completions Conference, Galveston, USA,

(2014c).

Growcock, F. B., Ellis, C. F. & Schmidt, D. D., “Electrical Stability, Emulsion Stability and

Wettability of Invert Oil-Based Muds”, SPE Drilling & Completion, pp. 39–46, (1994).

Haroun, M., Hassan, S., Ansari, A., Kindy, N., Sayed, N., Ali, B. & Sarma, H., “Smart Nano-

EOR Process for Abu Dhabi Carbonate Reservoirs.” Presented at Abu Dhabi International

Petroleum Exhibition and Conference, (2012).

Huang, T., Crews, J. B., Willingham, J. R., Pace, J. R. & Belcher, C. K., “Nano-Sized Particle-

Coated Proppants for Formation Fines Fixation in Proppant Packs,” US 7721803 B2,

(2010).

Husein, M. M. & Nassar, N. N., “Effect of Microemulsion on Copper Oxide Nanoparticles

Uptake by AOT Microemulsion”, Journal of Colloid and Interface Science, 316, 442–450,

(2007a).

Husein, M. M. & Nassar, N. N., “Study and Modeling of Iron Hydoxide Nanoparticles Uptake

by AOT (w/o) Microemulsions”, Langmuir, 23(26), 13093–13103, (2007b).

Page 87: A Field Application of Nanoparticles For Improved Downhole

75

Husein, M. M. & Nassar, N. N., “Nanoparticles Preparation Using the Single Microemulsion

Scheme”, Current Nanoscience, 4, 370–380, (2008).

ISO Standard 10416:2008, “Petroleum and Natural Gas Industries-Drilling Fluids-Laboratory

Testing”, International Organization for Standardization, (2014).

Jiao, D. & Sharma, M., “Mechanism of Cake Buildup in Crossflow Filtration of Colloidal

Suspensions”, Journal of Colloid and Interface Science, 162, 454–462, (1994).

Jihua, C. & Sui, G., “Rheological Behaviours of Bio-degradable Drilling Fluids in Horizontal

Drilling of Unconsolidated Coal Seams”, I. J. Information Technology and Computer

Science, 3, 1–7, (2011).

Jolivet, J.-P., Froidefond, C., Pottier, A., Chaneac, C., Cassaignon, S., Tronc, E. & Euzen, P. ,

“Size Tailoring of Oxide Nanoparticles by Precipitation in Aqueous Medium. A Semi-

Quantitative Modelling”, J. Mater. Chem., 14, 3281–3288, (2004).

Koutsoukos, P. & Kontoyannis, C. G., “Precipitation of Calcium Carbonate in Aqueous

Solutions”, Journal of the Chemical Society, Faraday Transactions., 1(80), 1181–1192,

(1984).

Midoux, N., Hoek, P., Paillers, L. & Authelin, J., “Micronization of Pharmaceutical Substances

in a Spiral Jet Mill”, Powder Technology, 104(2), 113–120, (1999).

Mukerjee, P. & Mysels, K. J., “Critical Micelle Concentrations of Aqueous Surfactant Systems”,

DTIC, 1–230, (1971).

Murty, B. S., Shankar, P., Raj, B., Rath, B. B. & Murday, J., “Textbook of Nanoscience and

Nanotechnology”, pp. 149–175, (2013).

Nassar, N. N., Hassan, A. & Pereira-Almao, P., “Application of Nanotechnology for Heavy Oil

Upgrading: Catalytic Steam Gasification/Cracking of Asphaltenes”, Energy Fuels, 25,

1566–1570, (2011).

Nelson, P. H., “Pore-Throat Sizes in Sandstones, Tight Sandstones, and Shales”, Association of

Petroeum Geologists Bulletin, 93, 329–340, (2009).

Nwaoji, C. O., Hareland, G., Husein, M. & Nygaard, R., “Wellbore Strengthening- Nano-Particle

Drilling Fluid Experimental Design Using Hydraulic Fracture Apparatus.” Presented at

SPE/IADC Drilling Conference and Exhibition, Amsterdam, Netherlands, (2013).

Oh, S. G., Jobalia, M. & Shah, D. O., “The Effect of Micellar Lifetime on the Droplet Size in

Emulsions”, Journal of Colloid and Interface Science, 156(2), 511–514, (1993).

Page 88: A Field Application of Nanoparticles For Improved Downhole

76

Parida, K. M., Pradhan, A. C., Das, J. & Sahu, N., “Synthesis and Characterization of Nano-

Sized Porous Gamma-Alumina by Control Precipitation Method”, Materials Chemistry and

Physics, 113(1), 244–248, (2009).

Patil, R. & Deshpande, A., “Use of Nanomaterials in Cementing Applications,” SPE 155607, pp.

12–14, (2012).

Rosen, M. J. & Kunjappu, J. T., “Surfactants and Interfacial Phenomena”, pp. 1–616, (2012).

Sensoy, T., Chenevert, M. E. & Sharma, M. M., “Minimizing Water Invasion in Shale Using

Nanoparticles.” Presented at SPE Annual Technical Conference and Exhibition, New

Orleans, USA, (2009).

Skauge, T., Hetland, S., Spildo, K. & Skauge, A., “Nano-Sized Particles for EOR.” Presented at

SPE Improved Oil Recovery Symposium, Tulsa, USA, (2010).

Taubert, A., Glasser, G. & Palms, D., “Kinetics and Particle Formation Mechanism of Zinc

Oxide Particles in Polymer-Controlled Precipitation from Aqueous Solution”, Langmuir,

18(11), 4488–4494, (2002).

Villas-Boas, M. B., Lomba, R. F. T., Sa, C. H. M., Oliveira, S. F. & Costa, J. F., “API Filtrate

and Drilling Fluid Invasion: Is There Any Correlation?” Presented at SPE Latin American

and Caribbean Petroleum Engineering Conference, Caracas, Venezuela, (1999).

Voorhees, P. W., “The Theory of Ostwald Ripening”, Journal of Statistical Physics, 38(1-2),

231–252, (1985).

Walz, J. Y., “Stability Behaviour of Clay/Nanoparticle Suspensions”, 56th Annual ACS Report

on Research, (2011).

Zakaria, M., Husein, M. & Hareland, G., “Novel Nanoparticle-Based Drilling Fluid with

Improved Characteristics.” Presented SPE International Oilfield Nanotechnology

Conference, Amsterdam, Netherlands, (2012).

Zamora, M., Broussard, P. N. & Stephens, M. P., “The Top 10 Mud-Related Concerns in

Deepwater Drilling Operations.” Presented at SPE International Petroleum Conference and

Exhibition, Villahermosa, Mexico, (2000).