You are on page 1of 102

REPORT JUNE

557 2016

Drilling waste management


technology review
A review of current and emerging techniques and technologies for
managing wastes arising from drilling activities in the oil and gas sector

manage care
Acknowledgements
Environment Committee

The International Association of Oil and Gas Producers (IOGP) appointed


AECOM to carry out a review of technologies and methods used for
managing waste generated by drilling activities in the oil and gas sector,
and this report presents the results of that review.

Photography used with permission courtesy of


©Gonzalez Thierry/TOTAL and ©Maersk Oil - Photographer Jacob
Russel (Front cover) ©ahopueo/iStockphoto (Back cover)

Disclaimer

Whilst every effort has been made to ensure the accuracy of the information
contained in this publication, neither IOGP nor any of its Members past present or
future warrants its accuracy or will, regardless of its or their negligence, assume
liability for any foreseeable or unforeseeable use made thereof, which liability is
hereby excluded. Consequently, such use is at the recipient’s own risk on the basis
that any use by the recipient constitutes agreement to the terms of this disclaimer.
The recipient is obliged to inform any subsequent recipient of such terms.

This publication is made available for information purposes and solely for the private
use of the user. IOGP will not directly or indirectly endorse, approve or accredit the
content of any course, event or otherwise where this publication will be reproduced.

Copyright notice

The contents of these pages are © International Association of Oil & Gas Producers.
Permission is given to reproduce this report in whole or in part provided (i) that
the copyright of IOGP and (ii) the sources are acknowledged. All other rights are
reserved. Any other use requires the prior written permission of IOGP.

These Terms and Conditions shall be governed by and construed in accordance


with the laws of England and Wales. Disputes arising here from shall be exclusively
subject to the jurisdiction of the courts of England and Wales.
REPORT JUNE
557 2016

Drilling waste management


technology review

Revision history

VERSION DATE AMENDMENTS

1.0 June 2016 First release


Drilling waste management technology review 4

Contents

Abbreviations 6

1. Scope 7
Methodology 8

2. Waste types 9
Drill cuttings 9
Drilling fluids 10
Waste cuttings 16
Cement 18
Interfacial mixtures 18
Spacer and completion fluids 19

3. Regulatory considerations 20
General 20
United States 21
Onshore 21
Offshore 21
Norway 24
United Kingdom 26

4. General considerations 27
Objectives of drilling waste management 27
Factors influencing choice of management routes 28
General 28
Reliability and throughput capacity 28
Performance 29
Cost 29
Environmental impacts 30
Portability 31

5. Management route – Drill cuttings 32


Collection and transport 33
Marine transport 33
Land transport 41
Solids control 41
Shale shakers 42
Hydrocyclones 46
Decanting centrifuges 48
Drilling waste management technology review 5

Secondary treatment 50
Mechanical cuttings dryers 50
Thermal treatment 53
Biological treatment 58
Physical and chemical treatment 62
Disposal 67
Offshore discharge 67
Re-injection 68
Onshore landfill 71
Salt cavern disposal 74
Application to land 75
Management of salt contaminated cuttings 78
Beneficial reuse 78
Construction materials 78
Fuel 79
Wetland restoration 80
Summary 80

6. Management routes – Interfacial mixtures and slops 86


General 86
Chemical treatment 87
Gravity separation 88
Dissolved air flotation 88
Decanting centrifuge 88
Disc separator 89
Filtration 89
Membrane separation 89
Ozone treatment 90
Microwave treatment 90
Electrocoagulation 90

7. Management routes – Other wastes 91


Drilling fluids 91

References 93
Drilling waste management technology review 6

Abbreviations

BFROC Base fluid retained on cuttings

CRI Cuttings reinjection

EMOBF Enhanced mineral oil base fluid

HOCNF Harmonized offshore chemical notification format

HPWBF High performance water based [drilling] fluid

LTMBF Low toxicity mineral oil base fluid

LTMO Low toxicity mineral oil

NABF Non-aqueous base fluid

NADF Non-aqueous drilling fluid

NF Nanofiltration

OBM Oil-based mud

SBF Synthetic based fluid

USEPA United States Environmental Protection Agency

VCD Vertical cuttings dryer

WBDF Water based drilling fluid


Drilling waste management technology review 7

1. Scope

This report provides a review of the drilling-specific wastes that are generated
during well construction activities, both onshore and offshore. This includes the
following main types of waste:
• drill cuttings and associated fluids, and
• interfacial mixtures.

The review focuses on waste types that are specific to drilling operations, or
which present significant challenges due to the quantities produced or potential
environmental impacts of these wastes.

This report focuses on technologies, methodologies, and processes for managing


waste once it has been generated.

It does not include consideration of well design or fluid selection, although


both of these are important in determining the quantity and type of drilling
waste generated, and hence should be considered as part of the overall waste
management plan for a drilling operation.

Waste types which are not included within the scope of this review include:
• water that is generated to the surface as a by-product along with the oil and
gas production (‘produced water’), and injection fluids returned to the surface
after fracture stimulation activities (‘flowback water’)
• general wastes (such as those generated by day-to-day activities of the
workforce), and
• potentially hazardous wastes likely to be generated in small quantities (such
as fluorescent light bulbs/tubes, batteries, and small-volume chemicals).

For each of those wastes, there is already a large amount of information in the
public domain about suitable technologies; the wastes in question are not unique
to drilling activities, and they are not generated in large quantities.
Drilling waste management technology review 8

Methodology
The methodology for the review is based on:
• Literature review of published sources of information. This takes into
account published papers in peer-reviewed journals, as well as conference
proceedings, presentations, technical reports and manufacturer’s literature.
• Stakeholder consultation. Telephone and face-to-face discussions were held
with a wide range of companies active in the drilling waste management
market, particularly technology vendors and oil field services companies,
as well as relevant departments in selected universities. IOGP member
organizations (representing operators) provided comments on initial drafts of
this report.

Information from these sources has been compiled into this report. The main body
of the report includes general descriptions of the types of technology, methods,
and processes that are applicable, with information about the constraints and
developments within each subject area.

Table 5 and Figure 39, located at the end of this report, provide summary
information for each technique:
• portability
• relative cost
• degree of commercialization
• qualitative environmental impacts
• advantages
• constraints
• achievable performance.
Drilling waste management technology review 9

2. Waste types

This section describes the waste types that are generated by drilling activities
within the scope of the review, and outlines their general characteristics and
the challenges these pose for waste management technologies.

Drill cuttings
The majority of wastes generated during drilling operations are spent drilling
fluids and drill cuttings. Drill cuttings are particles of crushed rock produced by
the action of the rotary drill bit as it penetrates into the earth. The drilling fluid is
pumped from the mud tanks on the rig, down the drill pipe, exiting through holes
in the drill bit, and returns to the surface via the annulus, which is the space
between the drill pipe and the drill casing or rock wall of the drilled hole (Figure 1).

Rotation of the drill bit at the bottom of the hole breaks off small chips of rock, called
cuttings, deepening the hole. The cuttings must be removed from the hole to the
surface in order to allow the drill bit to proceed. As the drilling fluid exits the drill bit,
it suspends the cuttings and carries them up the annulus to the surface where they
are separated from the drilling fluid by the solids control equipment on the drill rig.

drilling rig

mud pump

solids control
mud pits equipment

casing and
cement

open hole
drill pipe

annulus

drill bit drill collar

Figure 1: Drilling fluids circulating system of a drilling rig and well (From IPIECA/IOGP, 2009)
Drilling waste management technology review 10

Drilling fluids are normally reused until their properties become unsuitable for
the particular phase of the drilling operation. Drilling fluids used in offshore
operations may then be returned to shore for reprocessing or disposal by one of
the methods that are also used for disposal of fluids from onshore operations,
re-injected into an offshore well, or (in the case of water-based drilling fluids)
discharged to the sea if permitted by local regulations.

Processed drill cuttings may be discharged to the ocean, re-injected into a


suitable disposal formation, or transported to shore for treatment, disposal and/
or beneficial reuse. The choice of drilling waste management options depends
on the type of drilling fluids used, local regulations, drilling facility space/weight
limitations, environmental considerations, and cost–benefit analysis.

The physical and chemical characteristics of the drill cuttings will depend on the
formations drilled, and the type and quantity of any retained fluid. Drill cuttings
range in size from clay-sized particles (~0.002 mm) to coarse gravel (>30 mm)
and are irregular and angular. The chemical and mineral composition of cuttings
reflects that of the rock layers being penetrated by the drill. Cuttings from certain
formations (e.g. halite and other salts) can pose particular challenges for waste
management due to their chemical characteristics (Filippov, et al., 2009).

Drilling fluids
Drilling fluids, also referred to as ‘muds’, and are mixtures of fine-grained solids,
inorganic salts, and organic compounds dissolved or dispersed/suspended in
in a ‘continuous phase’ (the base fluid) which may be water or an organic liquid
continuous phase.

The two primary types of drilling fluid systems are: (IPIECA/IOGP, 2009)
• water-based systems (freshwater and saltwater systems). Water-based
drilling fluids (WBDFs) are the most widely used, and are generally less
expensive than other systems
• oil- or synthetic-based systems, collectively referred to as non-aqueous
drilling fluids (NADFs). NADFs have an oil or synthetic base fluid as the
continuous phase, and brine as the dispersed phase.

Water-based fluids
WBDFs are formulated mixtures of clays, natural and synthetic organic polymers,
mineral weighting agents, and other additives dissolved or suspended in fresh
water, seawater, brine, saturated brine, or a formate brine. The type of fluid
selected depends on anticipated well conditions or on the specific interval of the
well being drilled.
Drilling waste management technology review 11

WBDFs fall into two broad categories: non-dispersed and dispersed. They include
additives such as bentonite, polymers (including biodegradable polymers),
dispersants (e.g. lignosulfonates, lignitic additives, and tannins) and a variety of
other materials.

A large number of functional categories of additives are available for modifying


the physical/chemical properties of a WBDF to solve specific down-hole problems,
enabling it to function optimally during drilling of a well (Table 1).

Modern WBDFs rarely contain more than about ten of these additives, and most
are added in small amounts. The composition of the WBDFs may vary during
drilling of a single well because different additives may be required to drill different
well sections through varying geologic formations. Thus, the total inventory of
drilling fluid additives to drill all sections of a typical offshore well usually includes
about twenty additives, although the rig may also carry additional contingency
chemicals and additives.

Table 1: Functional categories of additives sometimes used in WBDFs to improve


drilling performance, with examples of chemical products in each category.

Category Example
Weighting materials Barite, calcium carbonate, ilmenite or hematite
Viscosifiers Clay, organic polymers
Filtrate reducers Starch, clay, lignite, polymers
pH control Inorganic acids and bases, most often caustic soda
Shale control Soluble salts such as potassium chloride (KCl), amines, glycols)
Lost circulation materials Inert insoluble solids such as calcium carbonate, ground nut shells,
graphite, mica and cellulose fibres
Lubricants Water-based lubricants, glycols and beads
Emulsifiers, surfactants detergents, soaps, organic fatty acids
Thinners lignite, lignosulfonates, polymers
Flocculants Inorganic salts, acrylamide polymers
Bactericides glutaraldehyde, triazine disinfectants
Pipe-freeing agents Water-based lubricants, enzymes, surfactants
Defoamers alcohols, silicones, aluminum stearate, alkyl phosphates
Calcium reducers Sodium carbonate, bicarbonate, polyphosphates
Corrosion inhibitors Amines, phosphates
Temperature stability acrylic or sulfonated polymers, lignite, lignosulfonate
Drilling waste management technology review 12

The most abundant additives (other than water at approximately 76 percent


by weight (% w/w)) in most WBDFs are weighting material (usually barite)
(approximately 14% w/w), inorganic salts (in several functional categories), and a
mud viscosifier (usually bentonite clay or a biologically derived organic polymer,
xanthan gum, guar gum) (approximately 6% w/w) (Figure 2).

other 4% gellants/other 1%
clay/polymer 6% brine 18%
non-aqueous
barite 14% fluid 46%

barite 33%

brine/water 76%

emulsifiers 2%

Water Based Drilling Fluid Non-Aqueous Based Drilling Fluid

Figure 2: Composition in weight percent of typical WBDFs and NADFs (IPIECA/IOGP, 2009).

High performance water-based fluids


High performance WBDFs (HPWBFs) have been developed because of strict
limitations on ocean discharge of NADF cuttings and the high cost of use and
disposal of NADF and NADF cuttings. HPWBFs are designed to replicate the
drilling performance characteristics of NADFs, contain non-toxic, biodegradable
ingredients, and minimize costs of drilling fluids and cuttings disposal (Marin, et
al., 2009).

Depending on the conditions at different depths in the well, special additives (often
different natural or synthetic polymers) in some of the categories listed in Table 1
can be added to the WBDF, for example to stabilize reactive shales, inhibit clay
swelling/sloughing, stabilize cuttings, increase the rate of penetration (ROP),
lubricate the drill string and bit, and stabilize mud viscosity at high temperatures
and pressures. However, despite great advances in HPWBM, NADFs still provides
superior technical performance in many cases, especially in providing chemical
shale stability, high rate-of-penetration, temperature stability and high lubricity.
Drilling waste management technology review 13

Non-aqueous drilling fluids


Conventional WBDFs often are not well suited for use in some drilling operations,
including the drilling of highly deviated, extended reach, and horizontal wells;
and deep, high pressure/high temperature wells often associated with offshore
developments. NADFs are typically used for drilling these wells.

NADFs are emulsions in which the continuous phase is a non-aqueous (water-


insoluble) organic base fluid (NABF), with water and chemicals as the internal
phase. NADFs provide better shale inhibition than WBDFs, and consequently
wellbore stability is maintained. NADFs are intrinsically lubricious; therefore the
ability to drill highly deviated extended reach and horizontal holes is enhanced
when compared with WBDFs.

In addition, NADFs are generally more stable in high-temperature applications,


such as those encountered in deep wells. Furthermore, NADFs are less susceptible
to the formation of gas hydrates that might potentially occur during deep-water
drilling operations (Canadian Association of Petroleum Producers, 2001).

NADFs can be formulated with diesel, mineral oil, or low-toxicity linear olefins,
paraffins, and esters. The olefins, paraffins and esters are often referred to as
‘synthetics’.

The ratio of the non-aqueous percentage to the water percentage in the liquid
phase of a non-aqueous based system is called its oil/water ratio. Non-aqueous
based systems generally function well with an oil/water ratio in the range of 65/35
to 95/5, but the most commonly observed range is from 70/30 to 80/20.

As with WBDFs, chemicals are added to the NADF; many of these additives provide
the same or similar functions to WBDF additives. Water, containing inorganic salt
and oil-soluble or oil dispersible additives, is dispersed into the non-aqueous
continuous phase and the resulting emulsion is stabilized with emulsifiers. NADF
viscosity is controlled by varying the amount of water dispersed into the non-
aqueous phase or by adding organophilic clays.

NADFs also contain thinners and weighting agents. A typical modern NADF
contains about 20–50% w/w NABF, 10–20% w/w brine, 0–50% w/w weighting agent
(e.g. barite), and <5% w/w other additives (Figure 2).

The environmental performance of a NADF is dictated by the base fluid used, and
the IOGP has grouped NABFs according to aromatic hydrocarbon concentrations
(a component that contributes to drilling fluid toxicity) as shown in Table 2.
Drilling waste management technology review 14

Table 2: IOGP classification of Non-Aqueous Base Fluids (International Association


of Oil and Gas Producers, 2003)

Category Properties
Group I: Non-aqueous These were the first NABFs used and include crude oil, diesel
fluids (high aromatic and conventional mineral oils. Diesel and mineral oils are refined
content) from crude oil and are complex mixtures of liquid hydrocarbons,
including paraffins, aromatic hydrocarbons, and polycyclic aromatic
hydrocarbons (PAHs). Group I NABFs are defined as containing more
than 5% by weight aromatic hydrocarbons, with PAH concentrations
greater than 0.35% w/w.

Because of concerns about toxicity and persistence in the ocean,


drill cuttings produced with Group I NADFs are no longer discharged
in most of the world. However, in situations where transportation of
cuttings to shore for disposal or reinjection of cuttings is possible,
such fluids may still be used offshore.
Group II: Non-aqueous These fluids, typically referred to as Low Toxicity Mineral Oil (LTMO)
fluids (medium aromatic or Low Toxicity Mineral Oil Base Fluids (LTMBF) were developed to
content) address concerns over the potential toxicity of diesel-based fluids.
Group II NABFs are also developed from refining crude oil, but the
distillation process is controlled to the extent that total aromatic
hydrocarbon concentrations (between 0.5 and 5%) are less than those
of Group I NABFs and PAH content is less than 0.35%, but greater
than 0.001%.
Group III: Non-aqueous This group of NABFs includes fluids produced by more extensive
fluids (low to negligible refining of a petroleum stock to produce enhanced mineral oil base
aromatic content) fluids (EMOBFs) or by synthesis of a specific, well-defined organic
fluid from non-petroleum precursors to produce synthetic base fluids
(SBFs). These fluids contain less than 0.5% w/w total aromatics and
less than 0.001% w/w total PAH. EMOBFs are produced by extensive
distillation, hydrogenation and separation of a refined petroleum
feedstock. The composition of EMOBFs depends on the feedstock and
the refining or separation processes used. SBFs are produced solely
by reaction of specific, well-defined chemical feed-stocks rather than
from a petroleum feedstock. By virtue of the source chemicals and
the manufacturing process, they contain no aromatic hydrocarbons.
The most frequently used synthetic hydrocarbons are esters,
polymerized olefins (linear alpha olefins (LAOs) and IOs (internal
olefins)), and synthetic branched and normal paraffins.

Pneumatic (air, mist, foam, gas) drilling fluids


Compressed air or gas can be used in place of drilling fluid to circulate cuttings
out of the wellbore. Pneumatic systems are implemented in areas where formation
pressures are relatively low and the risk of lost circulation or formation damage
is relatively high. The use of these systems requires specialized pressure-
management equipment to help prevent the development of hazardous conditions
when hydrocarbons are encountered.
Drilling waste management technology review 15

Four basic operations are included in this specialized category:


• dry air drilling, which involves injecting dry air or gas into the wellbore at
rates capable of achieving annular velocities that will remove cuttings
• mist drilling, which involves injecting a foaming agent into the air stream
that mixes with produced water and coats the cuttings to prevent mud rings,
allowing drill solids to be removed
• foam, which uses surfactants and possibly clays or polymers to form a high
carrying-capacity foam
• aerated fluids, which rely on mud with injected air to remove drilled solids
from the wellbore (World Oil, 2014).

Pneumatic-drilling operations require specialized equipment to help ensure safe


management of the cuttings and formation fluids that return to surface, as well as
tanks, compressors, lines, and valves associated with the gas used for drilling or
aerating the drilling fluid or foam. Common additives include (Rehm, et al., 2010):
• foaming agents:
–– ethoxyl alcohol ether sulphates
–– alpha olefin sulphonates;
• foam stabilizers/boosters:
–– amino propyl betaines
–– alkanol amides
–– sodium sulfosuccinates
–– alkyl-phenyl ethoxylates
• foam extenders:
–– xanthum gum polymer
–– hydroxyethylcellulose
–– carboxymethycellulose
–– bentonite.

Whilst foam and mist systems are mainly water-based, non-aqueous foam
systems are also available, using foaming agents such as silicone and
fluoraliphatic polymeric esters (Weatherford, 2012).
Drilling waste management technology review 16

Specialty products
Drilling-fluid service companies provide a wide range of additives that are
designed to prevent or mitigate costly well construction delays.

Examples of these products include:


• lost-circulation materials (LCM) help to prevent or stop downhole fluid
losses into weak or depleted formations. These may include sized calcium
carbonates, mica, fibrous material, or organic material such as crushed
walnut shells
• spotting fluids help to free stuck pipe
• lubricants for WBFs to ease torque and drag and facilitate drilling in high-
angle environments
• protective chemicals (e.g. scale and corrosion inhibitors, biocides, and H2S
scavengers prevent damage to tubulars and risks to personnel.

Waste cuttings
The drilling fluid circulates through the hole and returns to the surface as a
mixture of fluid and cuttings. Fluid is separated from the cuttings and recycled,
being recirculated back down the wellbore or being transferred/stored for use in
different holes. The cuttings are separated from the fluid by various means, and
become waste.

The fraction of solids removed will depend on the size of the particle, the efficiency
of the solids control equipment and the type of equipment (Figure 3).

Complete separation of the drilling fluid from the drilled solids is difficult to
achieve with conventional solids control equipment and a proportion of the fluid
is usually retained on the cuttings after the initial stage of separation (Figure
4). Removal of this fluid requires additional processing or cuttings cleaning
equipment.
Drilling waste management technology review 17

Colloids Silt Sand Gravel


(<2 μm) (2 – 74 μm) (74 – 2000 μm) (>2000 μm)
Decanting Centrifuge
removes >5 μm

4” Hydrocyclone (desilter)
removes >25 μm
% Solids Concentration

10” Hydrocyclone
(desander) removes >50 μm
Shale Shakers remove >74 μm

Fine drill solids


Drilled
and commercial clays Barite
Solids
(bentonite etc.)

0.1 1 10 100 1000

Particle Size (log10 Microns)

Figure 3: Particles sizes and solids removal equipment (Chevron)

Cuttings Treatment Solids Control


<0.1% Friction Derived thermal desorption 5% Centrifugal Cuttings Dryer 10% Shale Shaker 15%

1% 2% 3% 4% 5% 6% 7% 8% 9% 10% 11% 12% 13% 14%

% base fluid retained on cuttings (BFROC)


post-treatment (g-NABF/g-cuttings wet)

Figure 4: Cuttings treatment and residual base fluid on cuttings (Chevron)

The waste cuttings are therefore a mixture of the natural rock and soil material,
and the base fluid (which may be aqueous or non-aqueous), plus any associated
materials in the fluid, such as emulsifiers and brine salts, barite or calcium
carbonate and lost circulation materials (LCM). In some cases, there may also be
some reservoir fluids mixed with the cuttings.

This mixture of cuttings and fluids has the potential to cause impacts on the
surrounding environment and its disposal is therefore regulated.

The hydrocarbon content of the cuttings is referred to by a number of terms,


including OOC (oil on cuttings) and BFROC (base fluid retained on cuttings), which
will be the terminology used in this report. Retention of fluid on cuttings also
represents a financial loss, since new fluids must be purchased to replace those
disposed of as waste.
Drilling waste management technology review 18

Cement
Cement is used to hold casing in place and to prevent fluid migration from
subsurface formations.

Cementing is the process of mixing a slurry of cement, cement additives and water
and pumping it down through casing to critical points in the annulus around the
casing or in the open hole below the casing string.

The three principal functions of the cementing process are:


• to restrict fluid movement between the formations (zonal isolation)
• to bond and support the casing
• to ensure well integrity.

In addition to isolating oil, gas, and water-producing zones, cement also aids in:
• protecting the casing from corrosion
• preventing blowouts by quickly forming a seal
• protecting the casing from shock loads in deeper drilling
• sealing off zones of lost circulation or thief zones.

Cement circulated back to the surface during a cement job is referred to as


cement returns which become a waste stream considered in this review.

Interfacial mixtures
Interfacial mixtures are more commonly referred to as ‘slops’. Slop water or
slop mud is a waste stream which is produced when the drilling fluid becomes
contaminated with water (Ivan & Dixit, 2006); or an aqueous fluid (such as a
completion brine) becomes contaminated with NADF or reservoir hydrocarbons.
These waste streams arise from displacement interfaces (see below) and
operations such as cleaning the drill floor, shaker room, pump room, fluid tanks
and other areas where spillage can occur. Contamination of NADF by water affects
mud properties: it lowers the oil/water ratio (OWR), increasing viscosity, decreasing
emulsion stability and ultimately forming an unusable drilling fluid.

The composition of slops can be highly variable, and can include both oil-in-water
and water-in-oil emulsions.
Drilling waste management technology review 19

Spacer and completion fluids


Spacers are used primarily when changing fluid types and to separate fluid from
cement during cementing operations. In the former, an oil-base fluid must be kept
separate from a water-base fluid. In this case, the spacer may be base fluid. In the
latter operation, a chemically treated water spacer usually separates drilling fluid
from cement slurry.

Spacer fluid that is displaced to the surface will become waste if it cannot be
reused or if it is contaminated with cement or non-aqueous drilling fluid – in the
latter case, it will require being managed as slops, since it will be an interfacial
mixture of aqueous and non-aqueous liquids.

Completion fluid is placed in the well to facilitate final operations prior to initiation
of production, such as setting screens, production liners, packers, downhole
valves or shooting perforations into the producing zone. The fluid is used to control
a well, without damaging the producing formation or completion components,
should downhole hardware fail. Completion fluids are typically brines (chlorides,
bromides and formates), but in theory could be any fluid of proper density and
flow characteristics. The fluid should be chemically compatible with the reservoir
formation and fluids, and is typically filtered to a high degree to avoid introducing
solids to the near-wellbore area.
Drilling waste management technology review 20

3. Regulatory considerations

Whilst a detailed review of regulations governing drilling waste management is


outside the scope of this review, it is nevertheless important to understand the
general regulatory environment since many of the recent advances in drilling
waste treatment technologies have occurred in response to regulations.

General
The choice of drilling waste management techniques and the development
of technologies are tied closely to the regulatory environment, and hence the
following sections provide a limited discussion of relevant regulations in a small
selection of jurisdictions. The regulations applying to drilling waste management
vary greatly throughout the world and this section is not intended to present a
comprehensive overview of all applicable regulations.

For offshore operations, the most significant global difference in regulations is


between the approach adopted by the United States Environmental Protection
Agency (USEPA) in the Gulf of Mexico, and that adopted by the countries
bordering the north-eastern Atlantic that are contracting parties to the Oslo-Paris
Convention or OSPAR.

The USEPA approach is based on a determination that appropriately controlled


discharges represent a best-available technology after consideration of effects
on the marine environment and the non-water-quality impacts that would be
associated with a zero-discharge approach. The USEPA approach also takes into
account non-water quality impacts that would have been associated with a zero-
discharge approach, improvements in solids control equipment, and the results of
seabed studies of the impacts of NADF cuttings on the seafloor.

In contrast, OSPAR decision 2000/3 prohibited discharge of NADF fluids and NADF
cuttings containing >1% BFROC (Freidheim & Candler, 2008) in the North Sea.
The 1% BFROC limit was intended to prevent the discharge of any NADF adhering
to cuttings. Under the OSPAR system, discharges of WBDF, WBDF cuttings, and
other drilling wastes used offshore are regulated primarily by controls on chemical
substances used in the drilling process.
Drilling waste management technology review 21

United States

Onshore
Management of drilling wastes varies from one jurisdiction to another, ranging
from limited controls, to mandated, dedicated landfills. For example, California
allows, through CCR Title 27 §20090, “discharges of drilling mud and cuttings
from well-drilling operations, provided that such discharges are to on-site sumps1,
do not contain halogenated solvents, and at the end of drilling operations, the
discharger accomplishes either of the following tasks: (1) removes all wastes from
the sump; or (2) removes all free liquid from the sump and covers residual solid
and semisolid wastes, provided that representative sampling of the sump contents
after liquid removal shows residual solid wastes to be non-hazardous.”

Offshore

Discharge of chemicals (including drilling fluids)


The USEPA, Department of Energy (DOE), the former Minerals Management
Service, and numerous companies and industry associations worked together
using an innovative expedited rulemaking process to finalize new effluent
limitations guidelines (ELGs) for synthetic based fluids (SBF) in 2001. SBF is a
term used in USEPA parlance to indicate, effectively, NADF containing Group III
base fluids. Those rules allow for discharge of SBF cuttings, subject to various
restrictions but prohibit the discharge of SBFs themselves.

SBFs may not be discharged, except for small amounts adhering to cuttings and
certain small volume discharges. SBF cuttings discharges are allowed, subject to
several limitations. They must meet the same limits as WBDFs and cuttings for
free oil, cadmium and mercury in barite used to make up drilling fluids, and toxicity
of the suspended particulate phase. The base fluid used for drilling fluids must
meet limits for polynuclear aromatic hydrocarbon (PAH) content, sediment toxicity,
and biodegradation rate. In addition, the discharged material is subject to limits
on sediment and aquatic toxicity, base fluid retention on cuttings (6.9% for internal
olefins and 9.4% for esters), and formation oil.

1
Sumps are earthen excavations used to store drilling waste, including drill cuttings and cement returns.
Drilling waste management technology review 22

Oil-based muds (OBMs) in USPEA parlance are, effectively, NADFs containing


Group I and Group II base fluids, inverse emulsion fluids, oil contaminated fluids,
and fluids to which any diesel oil has been added. OBMs cannot be discharged
under USEPA regulations. Mineral oil may be used only as a carrier fluid
(transporter fluid), lubricity additive, or pill in water-based drilling fluids and may
be discharged with those drilling fluids, provided the discharge continues to meet
the no free oil and toxicity limits.

Effluent limits for drilling fluids and cuttings discharges on the U.S. Gulf of Mexico
Outer Continental Shelf based on USEPA permits GMG290000 (central and western
Gulf of Mexico) and GMG460000 (eastern Gulf of Mexico) are summarized in Table 3.

Table 3: Summary of effluent limits for drilling fluids and cuttings discharges on the
US Gulf of Mexico Outer Continental Shelf, based on USEPA permits GMG290000
(central and western Gulf of Mexico) and GMG460000 (eastern Gulf of Mexico)

Regulated parameter Discharge limitation/provision


Water-based drilling muds and cuttings
Drilling mud toxicity 30,000 parts per million (ppm) (daily minimum and a monthly
average minimum) (96 h LC50 of suspended particulate phase
with Mysidopsis bahia)
Cadmium in barite 3.0 mg/kg (dry wt)
Mercury in barite 1.0 mg/kg (dry wt)
Free oil No discharge (static sheen test)
Discharge rate 1,000 barrels per hour (bbl/h) maximum (does not apply to
drilling muds discharged prior to installation of the marine riser)
Discharges near Areas of Eastern Gulf of Mexico: No discharge of drilling muds and
Biological Concern cuttings from facilities within 1,000 m of an Area of Biological
Concern. Central and Western Gulf of Mexico: No discharge of
drilling muds within Areas of Biological Concern. Drilling mud
discharge rate within 544 m of Areas of Biological Concern is
limited based on distance and mud toxicity.
Discharges near ocean disposal Eastern Gulf of Mexico: No discharge within 1,000 m of a
sites Federally Designated Dredged Material Ocean Disposal Site
Synthetic-based muds (SBMs)
Discharges No discharge, except that which adheres to cuttings, small
volume discharges, and de minimis discharges. Small volume
discharges include displaced interfaces, accumulated solids in
sand traps, pit clean-out solids, and centrifuge discharges made
while changing mud weight. Allowable de minimis discharges
include wind-blown muds from the pipe rack and minor drips
and splatters around mud handling and solids control equipment
SBM cuttings
Drilling mud toxicity 30,000 ppm (daily minimum and a monthly average minimum)
(96 h LC 50 of suspended particulate phase using Mysidopsis
bahia)
Drilling waste management technology review 23

Regulated parameter Discharge limitation/provision


Cadmium in barite 3.0 mg/kg (dry wt)
Mercury in barite 1.0 mg/kg (dry wt)
Free oil No discharge (static sheen test)
Formation oil No discharge
Polynuclear aromatic 10 ppm PAH (as phenanthrene) in base fluid
hydrocarbon (PAH) content of
stock fluid
Sediment toxicity of base fluid 10-day LC50 from sediment toxicity test of the base fluid with
Leptocheirus plumulosus must not be less than the 10-day LC50
of the internal olefin or ester reference fluid.
Biodegradation rate of base Cumulative gas production of base fluid at 275 days must not be
fluid higher than that of the internal olefin reference fluid.
Base fluid retention on cuttings: 6.9 g/100 g of wet drill cuttings (end-of-well maximum weighted
C16–C18 internal olefin mass ratio averaged over all well sections)
Base fluid retention on cuttings: 9.4 g/100 g of wet drill cuttings (end-of-well maximum weighted
C12–C14 ester or C8 ester mass ratio averaged over all well sections)
Sediment toxicity ratio of 4-day LC50 of sample removed from solids control equipment
discharged drilling muds must not be less than that of the internal olefin or ester
reference drilling mud
Other Non-aqueous drilling muds
Oil-based drilling muds No discharge of drilling muds or associated cuttings
Oil-contaminated drilling muds No discharge of drilling muds or associated cuttings
Drilling muds to which diesel oil No discharge of drilling muds or associated cuttings
has been added
Mineral oil Mineral oil may be used only as a carrier fluid, lubricity additive,
or pill. Discharge allowed if it meets the limitations for toxicity
and free oil

The limits for drilling fluid retention on cuttings are 6.9% for internal olefins and
9.4% for esters. These are averages over all SBM well sections. Operators are
required to monitor the retained oil on cuttings (BFROC), by taking grab samples at
the solids control equipment once per day, or one sample for every 500 feet drilled
(up to three per day). When seafloor discharges are made during dual gradient
drilling, BFROC cannot be monitored and the USEPA specifies default values of
14% of base fluid retained on cuttings and 15% as the mass fraction of cuttings
discharged at the seafloor. The default values are to be averaged with results from
daily monitoring to determine compliance.

The BFROC limits were developed by the USEPA based on a statistical analysis
of data from 65 wells and representing four cuttings dryer technologies (vertical
and horizontal centrifuges, squeeze presses, and high-g-force linear shakers).
The upper 95th percentile of the BFROC data was used to set the BFROC limits.
(USEPA, 2000).
Drilling waste management technology review 24

Norway
Oil exploration in Norway occurs primarily along the Norwegian Continental Shelf.
This includes the areas where Norway exercises rights for economic development
under the United Nations Convention on the Law of the Sea and includes the North
Sea, the Norwegian Sea and a portion of the Barents Sea.

The key legislation that pertains to the disposal of drill fluids and cuttings in
Norway includes the following:
• Offshore Chemical Regulations, 2002
• Offshore Petroleum Activities Regulations, 2005
• Food and Environment Protection Act, 1985
• Convention on the Protection of the Marine Environment of the North East
Atlantic (OSPAR)
• OSPAR Decision 2000/3 on the use of Organic Phase Drilling Fluids (OPF) and
the Discharge of OPF-Contaminated Cuttings.

The Norway Climate and Pollution Agency regulates petroleum industry use of
drilling fluids, produced water, and chemicals, with water discharge permits. To
further protect marine waters, Norway first introduced a ‘zero discharge’ goal for
petroleum activities (Storting White Paper No. 58, 1996-1997). This goal was later
refined to mean zero discharge of environmentally hazardous substances, using
Best Available Techniques, and follows the precautionary principle (Storting White
Paper No. 25, 2002-2003). An advisory cooperative group composed of government
and industry representatives developed a common definition that identified
relevant technologies to achieve zero discharge, and created a standard manner to
report discharges. This advisory group found that a literal interpretation of the zero
discharge goal was not economically feasible or environmentally beneficial.

Norway defines zero environmentally harmful discharges as:


• zero discharge of all added environmentally hazardous chemicals classified
as ‘red’ or ‘black’ in the national classification system, and
• zero harmful discharges from natural compounds and chemicals classified
as ‘yellow’ or ‘green’.
Drilling waste management technology review 25

Norway established stricter requirements for drilling operations and produced


water in areas north of the 68th parallel in the Barents Sea and Lofoten area
(Storting White Paper No.38, 2003-2004). In summary, Norwegian discharge
requirements for drill fluids and cuttings are summarized in Table 4 (USEPA, 2011).

Table 4: Summary of Norwegian Drilling Fluid Regulations North and South of the
86th Parallel

Requirements Water-based drilling fluids and Drill cuttings associated with


associated drill cuttings non-aqueous (oil-based)
drilling fluids
Toxicity, biodegradation, For all components in chemicals used For all components in
bioaccumulation test offshore except ‘green’ chemicals list chemicals used offshore
requirements except ‘green’ chemicals list
South of the 68th Discharge permit for the chemicals Oil-based and synthetic-
Parallel in the drilling fluid is required. Only based cuttings restriction: no
‘green’ and ‘yellow’ chemicals allowed. discharge if base oil content
Discharge of fluids and cuttings allowed is >1%
from the entire well. Some limitations,
e.g. no discharge of solids containing
>1% oil, by weight. No toxicity testing
of used drilling fluids. Heavy metals in
barite as low as possible
North of the 68th Same requirements as south of 68th No discharge allowed
Parallel parallel. In addition, discharge of fluids
and cuttings allowed from only the
‘top hole’ section of the well, if the
discharge would not cause significant
environmental impacts. All other
cuttings or fluids must be re-injected or
barged to shore
Drilling waste management technology review 26

United Kingdom
The Offshore Chemical Notification Scheme (OCNS) manages chemical use and
discharge by the UK and Netherlands offshore petroleum industries. The OCNS
was originally introduced by the UK in 1979.

In 1993, the UK Government introduced a revised scheme, which classified


chemicals using test protocols approved by OSPAR. This was modified in detail in
early 1996 to meet the requirements of the OSPAR Harmonized Offshore Chemical
Notification Format (HOCNF), which coordinates the testing requirements for
oilfield chemicals throughout the NE Atlantic sector.

The OCNS uses the OSPAR Harmonized Mandatory Control Scheme (HMCS),
developed through the OSPAR Decision 2000/2, on a system for the use and
discharge of offshore chemicals (as amended by OSPAR Decision 2005/1) and
its supporting recommendations. This ranks chemical products according to
Hazard Quotient (HQ), calculated using the CHARM (Chemical Hazard and Risk
Management) model.

With respect specifically to drilling waste, the main driver for reductions in oily
discharges into the North Sea is the OSPAR Convention, which serves as the basis
for national laws governing discharge of drilling wastes in offshore waters of the
oil producing coastal states of Western Europe.

In accordance with the OSPAR Decision 2000/3:


• the use of diesel-oil-based drilling fluids is prohibited
• the discharge of whole organic-phase fluids2 to the maritime area is
prohibited
• the mixing of organic-phase fluids with cuttings for the purpose of disposal is
not acceptable
• the discharge into the sea of cuttings contaminated with organic-phase fluid
at a concentration greater than 1% by weight of dry cuttings is prohibited
• the use of organic-phase fluids in the upper part of the well is prohibited.
Exemptions may be granted by the national competent authority for
geological or safety reasons
• the discharge into the sea of cuttings contaminated with synthetic fluids shall
only be authorized in exceptional circumstances.

The OSPAR Decision 2000/3 virtually eliminated the discharge of NADF or cuttings
contaminated with these fluids unless the BFROC is <1%.

2
OSPAR terminology for non-aqueous fluids
Drilling waste management technology review 27

4. General considerations

This section discusses general considerations when evaluating drilling waste


treatment technologies.

Objectives of drilling waste management


Management of drilling waste should meet the two objectives of:
• compliance with environmental regulations, and
• ensuring that drilling operations are not unreasonably delayed.

In addition, many operators and regulators are also looking for drilling waste
management techniques that offer environmental benefits and are consistent with
the principles expressed in the waste hierarchy (Figure 5), but these considerations
can only be addressed providing the main objectives of regulatory compliance and
reliability/capacity are met.

The waste hierarchy is a general environmental principle, enshrined in regulations


such as the European Union Waste Framework Directive (European Union, 2008)
and which aims to improve the sustainability of waste management operations.
Under the Waste Framework Directive, waste legislation and policy of the EU
Member States are required to apply as a priority order the following waste
management hierarchy, which prioritizes waste prevention, re-use and recycling
over disposal, as shown in Figure 5.

PRODUCT (NON-WASTE) PREVENTION

WASTE PREPARING FOR RE-USE

RECYCLING

RECOVERY

DISPOSAL

Figure 5: The Waste Hierarchy (European Union figure)


Drilling waste management technology review 28

Factors influencing choice of management routes

General
There is a wide range of techniques and technologies available. No single option is
preferable in all cases. The development of a drilling waste management strategy
needs to take into account a range of factors in addition to the performance
characteristics of the technology itself.

These factors include:


• Type of fluid used. This is a complex interaction: fluid selection will influence
choice of waste management technique, but equally the available waste
management techniques may themselves be one of the factors which
influence the types of fluid that can be used (particularly in zero discharge
environments).
• Existing infrastructure. Some areas will have existing infrastructure
designed for (or capable) of managing drilling waste, in which case
transporting waste to these facilities may be the most cost-effective solution.
Conversely, in remote areas with little or no waste infrastructure, project- or
well-specific solutions may need to be developed. Offshore facilities may have
space and weight limitations that constrain the choice of waste management
technologies that can be used.
• Transport. The remoteness of well sites from one to another and from
existing infrastructure will help determine whether to use existing facilities,
or develop new facilities regionally or locally. Emissions and the risk of
accidental spillage during transport can be a significant determinant of
overall environmental impacts.
• Regulations. Any solution must be regulatory compliant. As well as
emissions limits as described in the previous section, some jurisdictions
have significant restrictions on the landfilling and beneficial reuse of waste
materials.

This section focuses on the characteristics of specific waste management


techniques and technologies. The main factors which need to be taken into
account when selecting a technology are discussed in the following sections.

Reliability and throughput capacity


Reliability and capacity can be considered together, since the overall ability of a
waste management system to deal with a given quantity of waste is a function of
these two factors.
Drilling waste management technology review 29

Drilling operations are extremely costly, and any slow-down or cessation of drilling
due to an inability to manage wastes will be prohibitively expensive, and as such it
is critical that the throughput capacity of the drilling waste management system
does not constrain the rate of penetration (ROP), which is a metric for the progress
of the drilling operation. This is a particular issue for offshore drilling: storage
space on a platform or mobile offshore drilling unit (MODU) is very limited and
hence the waste management system needs to be able to treat the waste at the
same rate that it arises. If the rate of waste treatment falls behind the rate of
drilling waste generation, a backlog of waste will result; storage space will rapidly
run out and drilling operations must be paused.

Cuttings processing rates can be optimized by combining the use of bulk storage
with cuttings cleaning equipment so that cleaning may take place at the same time
as well operations that do not generate drill cuttings are being carried out. This
allows for more efficient use of the cuttings cleaning equipment but does increase
the footprint of the overall package.

Having a high throughput capacity is of little use if the system is not reliable
and prone to breakdown. As soon as a system goes out of operation, waste will
immediately build up and drilling will be constrained.

Therefore, when evaluating treatment technologies, particularly for offshore


applications, operators will look carefully at both throughput capacity and reliability.
The need for reliability is one reason why simple options such as ‘skip and ship’,
i.e. transferring cuttings in boxes (skips) to support vessels and shipping them to
shore, remain popular, since it also offers high reliability and predictable availability.

Performance
Where a particular discharge limit or intended end-use is in place, it is of course
essential that the chosen treatment system is capable of meeting these limits.
Treatment technologies vary in the end-points they are able to achieve. In theory, it
is possible to develop a system which can meet almost any required environmental
standard but the system’s usefulness may be compromised by its cost, reliability,
and energy use, and its space and weight requirements.

Cost
The cost of waste management is of course an important determining factor; but
cost considerations are not as straightforward as simply comparing the Capex
and Opex of alternative systems against the waste disposal costs for the resulting
outputs. NADFs are expensive, but their use may reduce the time taken to drill
the well and being able to recover a higher proportion of NADF back into the fluid
system can be a significant factor in deciding which treatment option to use.
Drilling waste management technology review 30

As well as routine costs, the operator should also consider the potential for
additional costs that may arise due to the unavailability of the chosen system. For
onshore operations, this may simply be additional transport costs to an alternative
temporary storage location. The additional costs associated with system failure
offshore can be significant.

Environmental impacts
Environmental impacts of a product or process are usually assessed by means of
an Environmental Impact Assessment (EIA) at the project level. Developed countries
generally have well-defined EIA regulations and methodologies. International
standards are also available, such as those promulgated by the International Finance
Corporation (IFC). The IFC publishes Performance Standards and Environmental,
Health and Safety Guidelines which define IFC clients’ responsibilities for managing
their environmental and social risks and applies to all investment and advisory
clients whose projects go through IFC’s initial credit review process.

The EIA process comprises a number of stages, including:


• Scoping – determining the scope of activities and potential impacts that will
be assessed in the EIA
• Baseline assessment – identifying and quantifying the environmental
baseline of the potentially impacted area prior to the project commencing
and identifying sensitive receptors
• Impact assessment – assessing the impacts of the project on the
environment and particular sensitive receptors, sometimes by means of a
comparative assessment of different alternatives, and
• Mitigation – identifying mitigation measures that can be put in place to
reduce the environmental impacts.

In some jurisdictions, such as the European Union, there is a requirement for


operators to demonstrate that they are using the best available techniques (BAT)
for waste management, where:
• “Best” means most effective solution in achieving a high general level of
protection of the environment as a whole.
• “Available” means those developed on a scale which allow implementation
in the relevant industrial sector, under economically and technically viable
conditions, taking into consideration the costs and advantages, whether
or not the techniques are used or produced locally, as long as they are
reasonably accessible to the operator.
• “Techniques” includes both the technology used, and the way in which the
installation is designed, built, maintained, operated and decommissioned
(European Union, 2010).
Drilling waste management technology review 31

Assessing environmental impacts for a particular drilling waste management


technique is not straightforward and generally needs to be done on a project-
specific basis. The impacts will depend not just on direct impacts such as
emissions and resource consumptions from the technology itself, but also impacts
associated with transport, pre-treatment and fluid/chemical use, transport, and
final disposal.

Life Cycle Assessment (for a process or product) is another approach which may
also be used as a means of comparing different approaches to drilling waste
management in some cases. In a Life Cycle Assessment, the emissions and
resources consumed that can be attributed to a specific product are compiled and
documented in a Life Cycle Inventory. An impact assessment is then performed,
considering human health, the natural environment, and issues related to natural
resource use. Life Cycle environmental impacts of treatment processes can vary
considerably, and may be difficult to accurately determine (European Commission
Joint Research Centre, 2010).

Portability
The most significant distinction in the context of this report is between technologies
that can be used offshore, and those that are limited to onshore use. However,
even for onshore technologies there is a distinction between larger facilities, static
facilities, and smaller mobile facilities that can move between rig sites.

Technologies suitable for use offshore must generally be compact so as to fit


on-board rigs with very limited space, and also be capable of certification to the
various standards required for offshore use. In particular, offshore equipment may
need to be certified as suitable for use in potentially explosive atmospheres.

The installation of new offshore equipment may require structural rig modification
and/or upgrades such as increase of deck load capacity. This requires extensive
engineering studies which can delay the installation of such equipment as well as
adding significant costs on the project.

Equipment which can be modularized, containerized, or skid mounted can


generally be mobilized and demobilized to sites more rapidly and cheaply than that
which requires more complex on-site installation.
Drilling waste management technology review 32

5. Management route –
Drill cuttings

This section describes the main technologies and methods used for
managing drill cuttings.

Effective management of the wastes from drilling operations requires


consideration of many interrelated factors. Within this review, a distinction is
drawn between:
• solids control technologies, which are principally concerned with managing
the quality of the drilling fluid, minimizing overall the generated volume of
waste to deal with, and
• secondary treatment technologies, where the main aim is to further treat the
solids produced by the solids control systems to facilitate their management
or disposal.

Although this distinction is useful to categorize technologies, it should be


noted that these types of technologies cannot be considered in isolation: the
performance of the solids control system will influence which types of secondary
treatment technologies are used, and similarly the availability and effectiveness of
secondary treatment technologies will influence fluid selection and solids control
techniques.

A combination of different management options is chosen for each particular


project, depending on factors including fluid type, location (geographically and
onshore/offshore), proximity to existing facilities, regulatory requirements,
availability of suitable end-uses and financial considerations. The schematic
below (Figure 6) illustrates the various options that can be used for onshore
drilling or offshore drilling with a riser installed. It should be noted that no single
management option is used in isolation, and in most cases the management route
will entail the use of a range of different technologies in combination. In particular,
secondary treatment may or may not be required, depending on the intended final
disposal option for the cuttings.
Drilling waste management technology review 33

Re-Injection
Drilling
Mud System
Operation
Re-Injection
Slurrification
Well

Solids Control
Sea Disposal

Shale Shaker

Secondary
Land Disposal
Treatment
Desander /
desilter Landfill
Cuttings Dryer
Disposal

Construction
Centrifuge Thermal
Material / other
Treatment
Re-use

Biological Land
Treatment Spreading
Key
Cuttings + fluid
Fluid
Cuttings
Stabilization

Figure 6: Schematic showing possible management options for drill cuttings

Collection and transport


Once separated from the fluid system, drilling wastes require transportation to the
point of treatment or disposal. In some cases, the secondary treatment or disposal
point is remote from the rig site. Transportation of drilling wastes is potentially
costly and has both environmental and safety risks.

Marine transport
Transportation of drilling wastes offshore can be particularly problematic due to
the need for intermodal transfer and the difficulties of materials handling on board
a congested offshore drilling rig. However, if regulations do not allow discharge of
cuttings or fluids at the drill site, and the drilling rig cannot accommodate on-site
re-injection, transfer either to an offshore site suitable for discharge or re-injection
or to an onshore treatment and disposal site will be necessary.
Drilling waste management technology review 34

Cuttings handling
On the rig itself, cuttings need to be transferred from the solids control equipment
to containers or skips. This is commonly done using gravity (slides and chutes
– Figure 7), by screw conveyors (augers – Figure 8 and Figure 9), and can also
be done pneumatically using vacuum collection systems (Figure 10) or positive
pressure pneumatic conveyance.

Screw conveyors are very efficient at transferring cuttings and are one of the
most economical methods of handling bulk materials. They may comprise single
or multiple units and are often the best choice over short distances, but are less
efficient over longer distances or for complex installations. For example, screw
conveyors cannot bend around obstructions or easily convey cuttings up or down
(Morris & Seaton, 2006). Screw conveyors also represent a significant safety
hazard if used incorrectly or without the correct guards (International Association
of Drilling Contractors, 2010).

Pneumatic conveyance uses differential air pressure to convey cuttings from the
shale shakers through pipes and hoses to cuttings storage boxes, bulk storage
tanks (silos) or processing equipment. Both screw conveyor and pneumatic
conveyance systems collect, store and move drilled cuttings within an enclosed
environment, minimizing spills and maximizing containment.

Figure 7: Onshore cuttings chute and Figure 8: Onshore auger based cuttings
collection trough (Photo – Chevron) collection and transfer system (Photo – Chevron)

Figure 9: Offshore Swivel auger based Figure 10: Dual vacuum cuttings transport
cuttings collection system (Photo – Chevron) system (Photo – M-I SWACO)
Drilling waste management technology review 35

Pneumatic conveyance
Pneumatic conveyance of drill cuttings is broadly split into two types: dilute/lean
phase conveyance; and dense phase conveyance (Morris & Seaton, 2005).

In dilute or lean phase conveying, particles are fully suspended in the conveying
air and transported at low pressure and high velocity. Examples of dilute
phase conveyance include blowing of cement and barite (dilute phase pressure
conveying) and dilute phase vacuum conveyance. Vacuum collection systems are
able to collect, store and move drilled cuttings within an enclosed environment,
minimizing spills and contamination.

Vendors state that vacuum collection systems are effective in minimizing safety
and environmental risks associated with cuttings collection and containment,
as well as being relatively simple to mobilize and demobilize. They can transfer
cuttings vertically as well as horizontally, which offers flexibility in space
requirements (MI-SWACO, 2014a). Vacuum collection systems are capable of
handling wet or dry cuttings (Scomi Oiltools, undated) but cannot convey drill
cuttings as far, or to the same vertical height as dense phase cuttings blowers
which are able to convey cuttings over a distance of 100 metres and to a vertical
height of 50 metres.

Dense phase conveying can also be used to ‘blow’ cuttings from the rig to storage
tank or boxes on a supply vessel for transportation and onshore treatment/
disposal. Dense phase conveying doesn’t suspend the particles in the conveying air
but transports them as slugs of material at high pressure and low velocity

Material conveyed by this method is loaded into a pressure vessel (also called
a blow pot) and when the vessel is full, the inlet and vent valves are closed and
compressed air is fed into the vessel. The compressed air extrudes the material
from the pressure vessel into the conveying line and to the destination (Figure 11).

Figure 11: Typical cuttings blower operating cycle (M-I SWACO)


Drilling waste management technology review 36

Once the vessel and conveying line are empty, the compressed air is turned off and
the vessel is reloaded. This cycle continues until all of the materials required for
the process have been transferred (Nol-tec Systems, 2015).

Cuttings pumps
Positive displacement/hydraulic sludge pumps can also be used for the
bulk transfer of drill cuttings but some systems require additional fluid or
preconditioning of the cuttings for optimum performance. Correct design and
routing of pipework is essential for successful bulk transfer operations. Pipe runs
should be straight and either horizontal or vertical. Narrow radius pipe bends
should be avoided wherever possible.

Transport to shore
Transport of drilling waste to shore or to other offshore facilities can be carried
out either using individual cuttings boxes (skip-and-ship) as shown in Figure 12, or
by bulk transfer from rig-based storage tanks to a vessel fitted with bulk storage
tanks (Figure 13) and thence to an onshore treatment facility.

Figure 12: Lifting empty skips from a supply Figure 13: Low profile cuttings storage
vessel (Photo – Chevron) tanks and transfer hose used to convey
cuttings from the platform to the supply
vessel (Photo – NOV Brandt)
Drilling waste management technology review 37

Skip-and-ship
Skip-and-ship involves the filling of cuttings containers (generally referred to as
boxes or skips) with cuttings (Figure 14 and Figure 15), which are then transferred
to supply vessels by crane or transferring cuttings to vessels via a blower system.

Figure 14: Blowing Cuttings to Cuttings Figure 15: Pneumatic skip filling station
boxes (Photo – NOV Brandt) (Photo – Chevron)

Cuttings boxes (e.g. Figure 16) typically have a capacity of four to eight tonnes
(Morris & Seaton, 2006). Developments in cuttings boxes include:
• spring-assisted or lightweight aluminium lids for safer access
• sealed lids, reducing the risk of spillage
• high-capacity payloads, which provide a large capacity with a minimized
footprint
• stackable designs, which allow for increased capacity without requiring
increased deck space
• locking mechanisms offering improved security, and
• low height profile, reducing the danger of working at heights.

Figure 16: Low profile drill cuttings box (NOV Brandt)


Drilling waste management technology review 38

Cuttings boxes for offshore use need to be certified according to an international


standard such as EN 12079 and/or DNV 2.7-1. A new global standard for offshore
containers, ISO 10855, is expected to come into effect in 2015 (Lloyds Register
Energy, 2014).

The use of skip and ship has the advantage of being relatively simple but poses
environmental and safety risks, particularly relating to increased crane operations.
As noted by (Morris & Seaton, 2006), a typical offshore well can generate in excess
of 1,000 tonnes of cuttings and require several hundred cuttings boxes. These
boxes have to be lifted onto a boat, transported to the rig, lifted onto the rig, and
then lifted to the filling station on the rig. Once filled with cuttings, the box is lifted
from the filling station, transferred down onto the boat, and finally lifted off the
boat when it returns to the shore base (Figure 17) (Gilbert, et al., 2010). This means
six or more crane lifts are required for each cuttings box filled, and at 200 boxes
per well, this amounts to 1,200 individual crane lifts per well. This represents a
significant safety risk to workers at the rig site, on boats, and at the shore base.
Larger boxes may be used but this can be constrained by the available handling
equipment (Morris & Seaton, 2006).

Figure 17: Typical cuttings box logistics and life cycle (Gilbert, et al., 2010)
Drilling waste management technology review 39

Bulk transfer
Bulk transfer offers an alternative to skip-and-ship. Bulk transfer systems
eliminate the need for separate cuttings boxes, with cuttings being transferred
from the rig directly to bulk tanks on a vessel and then transferred out of these
tanks at the supply base. There are also situations where cuttings may be
transported in bulk to avoid discharge at an environmentally sensitive location.

Cuttings can be conveyed to and from bulk transfer tanks pneumatically (Figure
18 and Figure 19). The consistency of drilling wastes may vary from a slurry or
liquid, to a dry sand-like material, depending on the processing it has undergone.
Transporting dry cuttings requires less pressure (approximately 1–3 bar) and
hence less energy via blowers than wet cuttings which require approximately 7 bar.

The heterogeneous nature of drill cuttings can cause problems when emptying the
bulk tanks so a variety of designs have been developed to facilitate emptying.

This includes:
• single centre discharge point – this is a simple system, with few moving
parts that relies on the tank having a specially designed, high angle conical
bottom that establishes a mass flow of the cuttings when pneumatic
pressure is applied to the of the tank
• use of a multiple discharge point arrangement and a ‘spiked’ tank bottom
arrangement (Morris & Seaton, 2006)
• mechanically assisted discharge – this uses mechanical agitation to break
up any cuttings at the base of the tank – whilst effective, such systems are
more complex and costly, and
• cutting slurry tanks – slurrification can minimize the risk of clogging, but
increases the volume and liquid content of the waste to be discharged.
Drilling waste management technology review 40

Figure 18: Pneumatic skip filling station showing contingency bulk storage
tanks which can be used to provide contingency storage if weather restricts
skip handling operations (M-I SWACO)

Figure 19: Pneumatic (blower) based cuttings collection and storage system
(Photo – Halliburton BSS)
Drilling waste management technology review 41

Land transport
Onshore transportation of cuttings should be carried out in accordance with the
relevant regulatory requirements of the host jurisdictions. These may include the
use of transfer notes or manifests, display of placards or notices if wastes are
deemed dangerous or hazardous, or the use of suitable containers, which prevent
uncontrolled releases to the environment.

Transport of cuttings by road can give rise to large numbers of vehicle movements,
which in turn can increase the risk of road traffic accidents.

Solids control
The goal of modern solids control systems is to reduce overall well costs by
prompt, efficient removal of drilled solids while minimizing the loss of liquids (NOV
Brandt, 6th Edition) Since the size of drilled solids and cuttings varies greatly –
from cuttings larger than one inch in diameter to sub-micron size – several types of
equipment may be used depending upon the specific situation. Most solids control
systems include several pieces of equipment connected in series (Figure 20).

The first piece of equipment used to separate the solids from the mud is usually a
vibrating screen or series of screens, the shale shakers. These use a combination
of mechanical energy and sized screens to separate the drill cuttings from the
drilling fluid. Particles that are larger than the mesh openings are removed
but will always carry an adhered film of mud. The amount of mud retained/lost
on the cuttings can have a significant impact on the overall volume of waste
produced, the impact of the cuttings on the receiving environment and well costs.
Correctly functioning, high efficiency shale shakers play the primary role in waste
minimization and the greater the mechanical energy or the longer the retention
time of the cuttings on the screen the drier the cuttings and the lower the
concentration of fluid that adheres to the cuttings (NOV Brandt, 6th Edition). The
recovered drilling fluid may then be processed using hydrocyclones or centrifuges
to remove the finer solids that were not removed by the shakers.

Hydrocyclones with cone diameters of 6 to 12 inches are called desanders, and


hydrocyclones with cone diameters of less than 6 inches are called desilters. When
a shale shaker is used below hydrocyclones to ‘dry-out’ the cone’s discharge and to
minimize the loss of fluid, this combination is called a mud cleaner or mud conditioner.

Decanting centrifuges are commonly used to deweight drilling fluids (i.e. to remove
barite and drill solids) but may also be used to remove the fine silt and ultrafine
clay-sized solids. Two centrifuges may be used in series to treat weighted mud
(i.e. one that contains barite): the first to remove and possibly recover barite, the
second to remove fine solids and reclaim the liquid phase.
Drilling waste management technology review 42

Drill cuttings may also need secondary processing or cleaning prior to disposal
and this is discussed separately below.

Figure 20: Simplified solids control equipment sequence (Redrawn by Chevron from NOV
solids control Handbook NOV Brandt)

Shale shakers
Shale shakers are the first step in the drilling waste management process. 100% of
the mixture of fluid and cuttings returning from the hole passes through the shale
shaker and is separated into two fractions: one predominantly containing fluid, and
another predominantly containing cuttings. Both fractions may then be subject to
further treatment, either (in the case of fluids) to facilitate their continued use, or
(in the case of cuttings) to facilitate recycling or disposal.

Shale shakers are established technology and generally consist of a holding


chamber or ‘possum belly’ which receives the mixture of drilling fluid and cuttings;
a weir arrangement which distributes flow to the screens; the screens themselves
which are mounted within a basket. Below the screen basket is a collection pan
which collects the underflow and return it to the mud tanks for re-use (Figure 21).
Drilling waste management technology review 43

Figure 21: Key components of a shale shaker (Photo – M-I SWACO)

The motion of the screens separates the drilled solids from the drilling fluid (which
flows through the screens). The size of solids removed depends on the mesh of the
screens. The removed solids are retained on the screens discharged from the end
of the shakers to a holding area.

Screen selection depends on factors including: shaker design; drilling fluid


circulation rate, fluid properties, hole size, and penetration rate. Screen mesh
sizes are generally described by reference to the API 13C RP (ISO 13501) standard.
API 13C RP describes a methodology to define and compare the absolute
separation potential of any shale shaker screen with reference to an equivalent
standard ASTM test sieve.

The shaking motion of shale shaker baskets falls into one of the following
categories:
• unbalanced elliptical
• circular
• linear, or
• balanced or progressive elliptical.

Unbalanced elliptical and circular motion shaker systems are the least
complicated, relatively inexpensive and easy to maintain, but have limited capacity
to remove finer particles, and hence are often used as scalping shakers (i.e. the
initial shaker used to remove the coarsest solids).
Drilling waste management technology review 44

Linear and balanced elliptical motion shakers can use finer screens, provide
improved solids conveyance, and more efficient solid/liquid separation by improved
transfer of kinetic energy to solids. The ability of linear and balanced elliptical
motion shakers to convey solids upslope allows for the use of finer screens, and
hence removal of smaller particulars from the fluid.

Screen decks within baskets may be mounted horizontally, sloping either away
from the discharge end (inclined) or towards the discharge end (declined). The
deck angle can be fixed or adjustable and multiple decks may be installed within
the same unit (Figure 22).

Figure 22: Twin deck Shaker (NOV Brandt)

Shaker designs include cascade systems, in which scalping shaker with coarser
screens is mounted above a main shaker which has finer screens. These dual
systems can use either separate units, or separate screens within the same unit.
Cascade systems are particularly useful where solids loadings are high (e.g.
during rapid drilling of large diameter holes or where gumbo (a type of soft, sticky,
swelling clay) arrives at the surface) (ASME Shale Shaker Committee, 2005).

Recent advances in shaker design are driven by factors which include the need to
reduce space requirements, enhance the recovery rate of fluid from cuttings, and
improve automation.

The gravitational forces applied by shakers have increased from around 2 g’s, to
as high as 8–9 g’s in some cases. High g-force shakers are reported to provide the
capability to run finer screens, while producing drier solids from those screens
(Drilling Contractor, 2007). Recent developments also include enhanced capacity
with existing footprints, such as the use of multiple screens within the same unit.
Drilling waste management technology review 45

Advances in automation include automatic deck angle and g-force adjustment


in response to changing drilling conditions. G-force adjustment is made through
variable frequency drive (VFD) (Drilling Contractor, 2007). The introduction of
improved controlled acceleration techniques allows the shaker to continuously
adjust the shaker basket’s g-force to pre-programmed set points. This technology
is reported to allow for improved solids removal efficiency with the potential to
reduce rig non-productive time and improve rate of penetration (ROP) by helping to
yield a cleaner wellbore (El Dhorry & Dufilho, 2012).

In a comparative test between four conventional shakers and a vacuum-assisted


shaker, the retained oil-on-cuttings ranged from 3% to 8% (Aase, et al., 2013),
which was interpreted as demonstrating improvements in shaker design and
operation when compared on the reported BFROC’s of 5% to 15% achievable in the
1990’s. However, discussions with industry personnel indicate that BFROC in the
5–15% are still typical of shale shaker performance.

The test also noted the following key factors affecting shaker performance:
• screen wear and durability
• leakage rate
• ease of maintenance and operability

A review of vendors’ literature suggests that advances offered by the current


generation of shakers may include some or all of the following:
• ability to switch between motions (e.g. from balanced elliptical to progressive
elliptical) in response to changes in drilling conditions. This may allow
shakers to rapidly switch from modes suited to high- versus low-volumes of
fluid and solids loading, to a mode which optimizes the BFROC of the cuttings
• use of variable-frequency drives, maximizing control over the g-force of the
shaker
• ability to vary deck angles, allowing control of pool depth and beach length;
• improved screen durability
• multi-screen systems, which may include integrated scalping decks and
drying decks.

One specific technical advance in shaker design is the use of vacuum-assisted


technology to enhance performance. The MudCube™ system (Figure 23) uses a
combination of high air flow and vacuum, instead of the conventional g-force used
in vibrating shakers, to separate fluids from drilled solids (Kroken, et al., 2013).
Drilling waste management technology review 46

Figure 23: Vacuum Shaker (Photo Cubility)

The individual screens of a conventional shale shaker are replaced by a single


rotating screen conveyor belt, a continuous loop with set mesh size, which carries
drilling fluid and drilled solids forward, while air is pulled through this filter belt
taking with it the liquid phase (i.e. drilling fluid). An air knife and water knives provide
sheet-like, high intensity fluid streams that keep the filter belt from plugging. The
vendor states that the system can handle a normal flow rate of 270 m3/hour and is
capable of achieving less than 6.9% BFROC, and in some cases as low as 3%.

Trials of the system reported that advantages include: low noise level; no vapour
emissions; easy operation; and no direct contact of personnel with the fluid. The
elevated processing flow rate is reported to make the system attractive for offshore
use (Kroken, et al., 2013).

Other solutions offered by vendors include the use of a rotary brush and a vacuum
manifold which can be installed underneath a shaker screen or attached to the end
of a shaker box. The vendors state that use of the system results in a significant
decrease in BFROC and that the system has been successfully adopted in several
onshore locations in the US (Western Oilfield Equipment Ltd, 2013) (MI-SWACO, 2014).

Hydrocyclones
Hydrocyclones convert pressure generated by a centrifugal pump into centrifugal
force within the hydrocylcone cell, causing suspended solids to be separated from
the drilling fluid. In drilling operations, hydrocyclones use these centrifugal forces
to separate solids in the 15- to 80-micron range from the drilling fluid (ASME Shale
Shaker Committee, 2005). Solids-laden fluid is discharged from the lower apex of
the cone, and the cleaned drilling fluid is discharged as overflow (Figure 24).
Drilling waste management technology review 47

Most hydrocyclones are of a balanced design. A properly adjusted, balanced


hydrocyclone has a spray discharge of solids and a small amount of liquid at the
underflow outlet and exhibits a central air suction core. The apex opening relative
to the diameter of the vortex finder will determine the dryness of the discharged
solids.

Hydrocyclones are often referred to as desanders and desilters depending on the


size of particles they are intended to remove. Desanding units are designed to
separate sand-sized drilled and are used primarily to remove high solids volume
associated with fast drilling of large-diameter top holes. The desander, removes
sand-size and larger particles that pass through the shale shaker screens. Use of
desanders is generally discontinued when using more expensive polymer WBDF
or NADF. Use of desanders is generally not cost-effective with NADF because
they discharge a significant amount of the liquid phase (ASME Shale Shaker
Committee, 2005).

Desilter cones differ from desander cones only in dimensions and operate on
exactly the same principles. They are used to remove smaller particles than those
removed be desanders: drilled solids in the 12- to 40-micron range. They will also
separate barite particles in the 8- to 25-micron range.

Desanders may be used to reduce loading


downstream on desilters. Installing a
desander ahead of the desilter can relieve
a significant amount of solids loading
on the desilter and improves desilter
efficiency. Desanders remove a higher
mass (i.e., coarser drilled solids) during
periods of high solids loading; desilters
can then efficiently process the reduced
solids-content overflow of the desanders
(ASME Shale Shaker Committee, 2005).

Hydrocyclones should be designed to


provide maximum removal of solids with
minimum loss of liquid. Hydrocyclones
produce a wet discharge compared with
shale shakers and centrifuges (ASME
Shale Shaker Committee, 2005). Although
simple and robust in design, requiring low
levels of maintenance, hydrocyclones are
limited in their capability to remove very
fine or flocculated materials.
Figure 24: Hydrocyclone
process (NOV Brandt)
Drilling waste management technology review 48

Mud cleaners are a combination of hydrocyclones mounted above a shale shaker


dressed with find screens (Figure 25). Mud cleaners are designed to remove drilled
solids larger than barite and minimize mud loss. Mud cleaners are normally
positioned in the same location as desilters in a drilling-fluid system (ASME Shale
Shaker Committee, 2005).

Figure 25: Mud Cleaner (NOV Brandt)

Decanting centrifuges
Fine solids accumulate in drilling fluid either from drilling of softer sediments,
or as a result of coarse particles breaking up within the fluid as it circulates or
during the initial solids control processes. These fine particles can adversely affect
fluid properties and hence hole conditions: the only effective ways to reduce the
concentrations of these fine particles are dilution or centrifugation.

Decanting centrifuges are mechanical devices used for the separation of solids
from slurries by accelerated sedimentation. As the drilling fluid is passed through
a rapidly rotating bowl, centrifugal force moves the heavier particles to the bowl
wall, where they are scraped toward the underflow discharge ports by a concentric
auger, also called a screw, scroll or conveyor, which rotates at a slightly slower
rate than the bowl.
Drilling waste management technology review 49

The separation of the heavier particles divides the processed fluid into two
streams: the heavy phase, also called the underflow or cake; and the lighter
phase, which is called the overflow, light slurry, effluent, or centrate.

Tricanting centrifuges are also available, which can separate the inflow into three
separate streams, usually a solid and two liquid phases but these have limited
application in the treatment of drilling fluid wastes because of the high solids
content of drilling fluids.

Two different bowl designs are available: conical, in which the entire bowl is cone
shaped; and cylindrical/conical, in which the effluent (or light slurry) end of the
bowl is cylindrical (Figure 26). This configuration is preferred for drilling-fluids
applications because it offers greater capacity (ASME Shale Shaker Committee,
2005).

The elevated centrifugal forces created by the rotation of the bowl accelerate the
sedimentation process so that separation that might take hours or days under the
normal gravitational force of 1 g in an undisturbed container is achieved in seconds
at the 400–3000 g generated by the centrifuge.

Figure 26: Decanting Centrifuge with the bowl cover open (Derrick Equipment)

Centrifuges can be used to recover fluid from hydrocyclone underflows, thereby


reducing drilling waste volume. The process returns the finest solids to the
fluid system, which can degrade drilling fluid performance (ASME Shale Shaker
Committee, 2005).
Drilling waste management technology review 50

Whilst most centrifuges use hydraulic or variable-frequency drives, some vendors


offer centrifuges which use magnetic couplings. These couplings use an induced
magnetic field to transfer rotational energy from the motor to the rotor. The
strength of the magnetic field depends on the gap between the magnets on the
input and output shafts, meaning that the coupling force can be varied as this gap
is varied. Vendors state that the use of magnetic couplings reduces vibration and
noise, extends equipment life and reduces maintenance costs (MI-SWACO, 2010).

A growing trend for onshore drilling is reported to be the use of high-volume, or


‘big bowl’, centrifuges that can remove 2–8 tons of solids per hour from the drilling
fluid and process more volume (Drilling Contractor, 2014).

Secondary treatment
Secondary treatment technologies are employed primarily to reduce the amount of
drilling fluid left on cuttings and thus saving costs by increasing recovery of drilling
fluids and reducing environmental impacts by reducing the concentrations of base
fluids on cuttings.

Mechanical cuttings dryers

Centrifugal cuttings dryers


Centrifugal cuttings dryers are used when drilling with NADF to further separate
fluids from the cuttings produced by shale shakers, which improves recovery of
the drilling fluid and reduces the BFROC concentration of the cuttings. Dryers
can be used either to meet a required BFROC content prior to discharge to the
environment, or as preparation for a subsequent form of treatment such as bio-
remediation.

Cuttings dryers are available with either vertical (Figure 27) or horizontal baskets
(Figure 28). Both configurations use a rapidly rotating cone within a wire basket
screen to exert a high g-force (400 g or more) on the cuttings, thereby separating
solids from liquids.

Vertical cuttings dryers (VCD) were developed from those originally used in the coal
mining industry, are designed to process large volumes of materials while being
able to withstand abrasive materials (Cannon & Martin, 2001). VCDs are available
in capacities of up to 60 tonnes/hour (Halliburton, 2014). The effectiveness of VCDs
in reducing the BFROC of cuttings depends on a number of factors, including
composition, viscosity, and temperature of the drilling fluid; the nature of the
formation being drilled; and the particle size of the cuttings. Reductions of BFROC
from 11.47% to 3.99% are reported (Cannon & Martin, 2001).
Drilling waste management technology review 51

Suppliers of VCD systems quote in their product literature that they are
typically able to achieve BFROC levels of <5% (Halliburton, 2014), and in some
circumstances as low as 1% for salt cuttings. (Weatherford, 2014). Suppliers also
highlight the economic benefits of increased recovery of drilling fluids. However,
fluids recovered using a dryer are likely to contain high levels of low gravity solids
and might therefore require further treatment (e.g. by high-speed centrifuge)
before reuse. VCDs are commercially proven in both onshore and offshore
applications.

Figure 27: Vertical cuttings dryer (NOV Brandt)

Figure 28: Horizontal cuttings dryer (NOV Brandt)


Drilling waste management technology review 52

Figure 29: Vertical cuttings dryer and decanting centrifuge installation (M-I SWACO)

Mobile cuttings dryers are available. They typically comprise a VCD and a high
capacity, decanting centrifuge positioned alongside a cuttings container (Figure 29).
Dried cuttings are collected in a cuttings bin where they are mixed with discarded
solids from the centrifuge. The dried cuttings are then available for transport and
final disposal. Effluent recovered by the dryer is collected in a holding tank before
being processed through the high speed centrifuge in order to be cleaned and
returned to the rig pit or active fluid system.

The advantages claimed for this system are: reduced stand-by time, a trailer-
mounted configuration offering reduced set-up time and costs; self-contained
power source eliminating reliance on rig power; low footprint and average BFROC
of 3–6.9%, reducing the total waste volume to be transported and disposed
(Halliburton, 2014).

Vacuum dryers
Other systems have been developed which use various combinations of pressure
differentials and high velocity air to enhance the separation of fluid from cuttings.
The Rotovac® rotary vacuum dryer (RVD) system, which is attached to the
discharge end of shale shakers, is reported by the supplier to be able to achieve
4% BFROC, and is most effective on larger size cuttings. The supplier also states
that the RVD, being installed adjacent to the shale shakers, reduces the need for
transfer of cuttings to a separate VCD (Halliburton, 2007).
Drilling waste management technology review 53

Thermal treatment
Thermal treatment can be used for managing certain types of drilling waste which
are contaminated by hydrocarbons: both to separate hydrocarbons from the matrix
material, but also for the destruction of hydrocarbons or other organic materials.

Incineration or thermal destruction involves the combustion of the hydrocarbon


components of the waste by incineration or thermal oxidation and may be by means
of dedicated incinerators or by co-combustion in a power plant or cement kiln.

Thermal separation works on the basis of heating a waste sufficiently to volatilize


the hydrocarbons that are absorbed into the matrix of the waste. Thermal
separation is commonly referred to as ‘thermal desorption’.

Thermal treatment is most commonly used for treating either oily sludges or drill
cuttings. These applications are quite distinct, in that oily sludges will contain a
high proportion of hydrocarbons and have a high calorific value, whereas cuttings
will usually have much lower hydrocarbon content (e.g. less than 10%). In some
circumstances the hydrocarbons may be desorbed and then used as a fuel or be
re-refined or be separately incinerated/thermally oxidized. However, either process
can be used independently. In many cases, thermal desorption is used to recover a
valuable hydrocarbon component of the waste stream and it is therefore important
to avoid combustion or degradation of the hydrocarbon fraction.

Incineration
Incineration is a widely adopted technique for managing many different types of
hazardous and non-hazardous wastes. Incinerators can be stand-alone facilities,
dedicated for burning waste; or co-combustion facilities which use waste as one of
their feed-stocks. A common form of co-combustion is the use of cement kilns for
waste incineration.

Different types of waste incinerators include (DEFRA, 2007):


• Grate systems: These include the moving grate furnace system, the most
commonly used combustion system for high through-put municipal waste
incineration. The waste is slowly propelled through the combustion chamber
(furnace) by a mechanically actuated grate. Waste continuously enters one
end of the furnace and ash is continuously discharged at the other. The
system is configured to enable complete combustion as the waste moves
through the furnace. Process conditions are controlled to optimize the waste
combustion, to ensure complete combustion of the feed. The end of the grate
normally passes the hot ash to a quench to rapidly cool the remaining non-
combustibles. Fixed grate systems are usually smaller and more commonly
used for remote locations and mobile plant. They typically comprise one or
more fixed chambers with the waste being moved by a series of rams.
Drilling waste management technology review 54

• Fluidized bed: Combustion is normally a single stage process and consists


of a lined chamber with a granular bubbling bed of an inert material such
as coarse sand/silica or similar bed medium. The bed is ‘fluidized’ by air
(which may be diluted with recycled flue gas) being blown vertically through
the material at a high flow rate. Wastes are mobilized by the action of this
fluidized bed of particles. The use of fluidized bed technology for incineration
is relatively limited, although it is widely applied to sewage sludge.
• Rotary kiln: Incineration in a rotary kiln is normally a two stage process
consisting of a kiln and separate secondary combustion chamber. The kiln is
the primary combustion chamber and is inclined downwards from the feed
entry point. The rotation moves the waste through the kiln with a tumbling
action which exposes the waste to heat and oxygen. Rotary kilns are often
used for hazardous waste incineration.

Waste incinerators have the potential to impact air quality and require careful
control of combustion conditions, often including the use of air pollution control
equipment such as filters and scrubbers. The outputs from an incinerator are air
pollution control residues (sometimes referred to as ‘fly ash’) and bottom ash or
clinker.

Waste water from the flue gas cleaning system can be treated in a separate
wastewater treatment plant if there is no direct access to a municipal waste water
treatment plant.

Residues are typically disposed of as follows:


• clinker – direct to secure landfill
• fly ash – direct to secure landfill
• waste water – to treatment plant.

Thermal desorption
Thermal desorption is primarily used to separate hydrocarbons from cuttings
drilled with NADF, thereby achieving the twin objectives of recovering valuable
hydrocarbons and lowering the BFROC value of the cuttings.

Reducing the BFROC of cuttings in turn allows them to be managed with much
greater flexibility and at lower cost. Provided the BFROC can be sufficiently
reduced, it may be possible to reuse cuttings (for example as construction fill) or
dispose of cuttings by marine discharge.

The factors to be taken into account when selecting and using a thermal
desorption process include (Stephenson, et al., 2004):
Drilling waste management technology review 55

• Regulatory compliance – the unit will need to comply with local


environmental regulations, including discharge requirements and permitting.
• Fluid recovery – the ability to recover costly drilling fluids in a state suitable
for reuse and without degradation of the fluid can be critical when evaluating
the cost-effectiveness of thermal desorption.
• Size and weight – equipment must be consistent with the available space and
weight constraints when used in offshore rigs.
• Flow rate and turndown – the unit should have sufficient flow rate to treat
the likely flow of cuttings, particularly offshore where storage is limited, and
may also need to cope with a range of flow rates.
• Treatment end points – the chosen process must be able to meet the
required standards for disposal or reuse.
• Standards – in particular, they must have the ability to meet the stricter
safety standards for offshore operations.

Several types of thermal desorption have been developed over the years, with their
initial application being in the field of contaminated soil treatment.

Key design factors for thermal desorption plant are:


• efficient transfer of heat to the waste matrix
• minimizing degradation of hydrocarbons.

Thermal desorption can be direct (where combustion is used to generate heat in the
same chamber as the desorption) or indirect (where heat is generated separately from
the desorption chamber) (Environment Agency, n.d.) or based on mechanical friction.
Most thermal desorption systems used in the oil and gas sector are indirect or friction
based systems. The use of direct units entails destruction of the recovered oil.

The main types of technology used for onshore thermal desorption are:
• Drum-type indirect thermal desorption units. These use a rotating drum
that is externally heated by burners and sealed to prevent the formation of
explosive mixtures. The drum is tilted; the cuttings are fed into the highest
end for treatment, and then discharged at the lower end. The vapour is fed
through a two-stage condenser to produce separate oil and water streams.
• Screw-type desorption units. These use a hollow screw or auger which
rotates and transfers cuttings from a feed hopper into a desorption chamber.
Heated oil is circulated through the screw, and/or through a jacket which
surrounds the desorption chamber, and the heat transferred to the cuttings.
Vapours are condensed in the same way as for a drum-type unit (Stephenson,
et al., 2004).
• Friction based thermomechanical desorption units. This is a thermal
desorption system which generates heat by friction, between the drill cuttings
and rotating arms within the desorption chamber.
Drilling waste management technology review 56

Drum-type indirect thermal desorption can be relatively inefficient and heat


transfer can be hard to control. The systems which use a fluid to transfer heat
from an external heat source to the interior of the chamber are more efficient and
controllable (Figure 30). Some systems also use electrical heating of the waste.

Figure 30: Indirect thermal desorption unit (NOV Brandt)

Figure 31: Low temperature thermal desorption unit (NOV Brandt)


Drilling waste management technology review 57

Most thermal desorption processes are fixed or mobile onshore installations


(Stephenson, et al., 2004) as illustrated in Figure 31. However, alternative types of
compact equipment are safe enough to be considered for offshore use.

One such example is the ‘Thermomechanical Cuttings Cleaner’ (TCC®) which has
been licensed to a number of firms (Figure 32). This is a thermal desorption system
which generates heat by friction, generated by the rotation of weighted arms within
the desorption chamber. The rotating arms pulverize the cuttings and the resulting
friction heats them to a high temperature (240 – 300 °C) which volatilizes the water
and hydrocarbons and leaves a clean dry powder (d90 <30 µm) for disposal. The
resulting system is compact enough to be installed on some offshore rigs.

Figure 32: Schematic of a thermomechanical desorption system (Halliburton image)

Performance of a thermal desorption system depends on a number of factors,


including:
• oil, water and solids ratios of incoming cuttings
• drive unit power
• mill chamber temperature.

A well-designed and operated thermal desorption process should be capable


of achieving BFROC of less than 1% (Stephenson, et al., 2004). In many cases
values much lower than this are achievable (MI-SWACO, 2011). A Norwegian study
reported an average BFROC of 0.03% for OBM cuttings treated using the TCC
process (Aquateam COWI, 2014).
Drilling waste management technology review 58

After BFROC, the other key performance consideration is the quality of the
recovered drilling fluid. Limiting the temperature and heating period can help
avoid degradation of fluids (Aquateam COWI, 2014).

The ability to recover fluids (and hence reduce the expenditure on fluids) is an
important part of the cost–benefit analysis when considering the use of thermal
desorption. Fluid degradation can occur through thermal cracking. Longer-
chain molecules undergo thermal cracking at lower temperatures than smaller
molecules (cracking can occur as low as 343ºC for C20 to C30 molecules). Thermal
cracking can create aromatics and other hydrocarbons which may adversely affect
the toxicity of the fluid and affect fluid performance (Seaton, et al., 2006).

The outputs from a thermal desorption facility are usually:


• recovered oil, which is usually recycled, but can be combusted
• recovered water, which is usually discharged to sewer or sea if permitted by
local discharge regulations
• dry cuttings, which can be landfilled, discharged to sea, or used as fill or
construction materials.

Research has been carried out into the use of microwaves for thermal desorption,
and studies carried out on a pilot scale (Pereira, et al., 2013), (Robinson, et al.,
2010). The technology is based on using microwaves to selectively transfer heat
to the aqueous fraction of the cuttings; as the water is heated and evaporates, it
also strips out the hydrocarbon phase from the solid cuttings matrix. These pilot
studies have indicated that a BFROC of 0.1% can be achieved (Robinson, et al.,
2010). Although academic-service company collaborations are working towards
commercialization of microwave desorption systems, as of mid-2015 such systems
are not commercially available.

Biological treatment
Composting is a controlled biological treatment process whereby organic
substances are converted by microorganisms to innocuous, stabilized by-products
(Paton & Fletcher, 2008). These by-products will typically be humus and microbial
biomass (McCosh & Getliff, 2004). It differs from land-farming in that composting
takes place under controlled conditions, either in the open air or within a vessel;
rather than by spreading directly onto the land surface and typically has a smaller
footprint than land farming (Figure 33).
Drilling waste management technology review 59

Figure 33: Biological treatment facility (Total photo)

• Aerated (turned) windrow composting – waste is formed into rows of long


piles called ‘windrows’ and aerated by turning the pile periodically by either
manual or mechanical means. The pile should be large enough to generate
sufficient heat and maintain temperatures, yet small enough to allow oxygen
to flow to the windrow’s core (US Environmental Protection Agency, 2015).
• Aerated static pile composting – waste is mixed together in one large pile
instead of rows. To aerate the pile, layers of loosely piled bulking agents are
added so that air can pass from the bottom to the top of the pile. The piles
can also be placed over a network of pipes that deliver air into or draw air out
of the pile (US Environmental Protection Agency, 2015). These are sometimes
referred to as biopiles.
• In-vessel composting – waste is fed into a drum, silo, concrete-lined trench,
or similar containment element, where the environmental conditions-
including temperature, moisture, and aeration-are closely controlled. The
apparatus usually has a mechanism to turn or agitate the material for proper
aeration. In-vessel composters vary in size and capacity (US Environmental
Protection Agency, 2015).

The composting process generates metabolic heat, and the rate of composting is
temperature-dependent: cooling or insulation may be required in order to control
the temperature of the waste (McCosh & Getliff, 2004). The heat generated within
the composting process enhances the rates of biodegradation and volatilization
when compared to land-spreading or land-farming, making it particularly suitable
in cold climates.
Drilling waste management technology review 60

When comparing composting with landfarming or other land application,


composting is recommended when:
• land space is limited
• the growing season is less than three to four months
• the soil condition is poor and does not support vegetation well
• large volumes of waste need to be treated in a relatively short period of time
• the salt content of the waste is compatible with eventual land application of
the compost
• BFROC concentrations are high (>20%) and control of volatiles is required,
and
• regulations prevent land spreading/land treatment or regulators have
imposed time constraints on treatment (McMillen, et al., 2002)

Successful composting of drill cuttings generally requires blending of the cuttings


with organic materials to provide the appropriate proportions of carbon, nitrogen,
and moisture, all of which are required for the composting process (McCosh
& Getliff, 2004). The composition of the drilling fluid has a significant effect on
the composition of the resulting composted product, and also on the reaction
rate of the composting process, with aromatic and branched-chain molecules
generally being more resistant to biodegradation than aliphatic and straight-chain
molecules of similar molecular weight, and lower molecular weight compounds
biodegrading faster (McCosh & Getliff, 2004). Salt and heavy metals (from brine
and weighting fluids) can also have an impact on the quality of the compost, and its
suitability for use as a soil improver.

Organic materials used for blending with cuttings include sawdust, wood chips,
straw and chicken manure (McCosh & Getliff, 2004). Experiments carried out to
determine the effect of base fluid type on composting rates suggested that (in
order of ease of degradation), linear paraffin degrades fastest, followed by linear
alpha olefin, then enhanced/low-toxicity mineral oil, followed lastly by diesel. In all
cases, it was possible to achieve total extractable hydrocarbon concentrations of
less than 0.1% after 112 days (McCosh & Getliff, 2004). Some fluid suppliers offer
fluids which are optimized for subsequent bioremediation or composting of the
cuttings.

The carbon:nitrogen (C:N) ratio is a critical parameter and a ratio of around 30:1 is
recommended for successful composting; other critical parameters are moisture
content (50% is reported as being optimum) and the ratio of cuttings to organic
matter (1:1 reported as being optimum) (McCosh & Getliff, 2004).
Drilling waste management technology review 61

Composting can be used as part of a combined treatment process. For example,


a case study from Oman reported the treatment of SBM drill cuttings using a
combination of chemical pre-treatment (using a proprietary free radical mixture),
followed by composting with hay and straw as organic matter. The case study
indicated that petroleum hydrocarbon concentrations in the cuttings were
reduced from 22% to less than 0.5% within 24 weeks (Nahmad, et al., 2014). Other
studies report cuttings being washed prior to composting (Yan, et al., 2011). Co-
composting field trials of C14 linear alpha olefin drill cuttings in windrows with
wood fiber (at a ratio of 1:3 cuttings: fiber) and fertilizer at two sites in British
Columbia (Canada) showed that the total extractable hydrocarbon concentration
reduced from 8.1 and 15.2% to 0.4 and 0.6% respectively, when composted over a
22 month period (Visser, et al., 2004).

In some cases, the concentrations of microorganisms can be artificially


augmented by adding an inoculum, if there is considered to be a deficiency of
naturally-occurring suitable organisms.

It is reported that nutrients in the form of nitrogen and phosphorous may be


needed to enhance the biodegradation rate of hydrocarbons in soil. Addition of
the necessary nutrients can readily be accomplished by adding agricultural-grade
fertilizers to supply nitrogen and phosphorous (McMillen, et al., 2002).

Vermicomposting is a particular process which uses worms to treat organic


waste. This approach was originally used in New Zealand to treat drill cuttings
(Getliff, et al., 2002) and was further investigated under controlled conditions in
Norway. However, this method does not appear to have been commercialized on
a widespread basis: one significant difficulty is the need to maintain over time
the healthy colonies of worms that are necessary for the composting process,
particularly when the supply of cuttings can be irregular or limited to a brief period.

In summary, composting is used for treatment of drilling waste in many


jurisdictions (e.g. UK, Nigeria, Canada, Australia and USA) but is only viable for
onshore treatment, either of waste generated onshore or offshore waste that has
been shipped to shore. The process can be relatively time-consuming and requires
relatively large treatment facilities; but can be particularly suited to remote
environments where other treatment infrastructure is lacking (Visser, et al., 2004).
Since the volume and mass of the waste is increased by addition of the bulking
agent, substantially more waste may require disposal unless the treated material
can be beneficially reused (e.g. as a growing medium or soil improver). The high
salt content of some drill cuttings can be a barrier to beneficial reuse.
Drilling waste management technology review 62

Physical and chemical treatment

Solidification and stabilization


Solidification/stabilization techniques use physical or chemical immobilization, to
transform drilling waste to a less hazardous form (Ball, et al., 2011).

Aims of stabilization and/or solidification include:


• to create products from waste, e.g. construction products (in solid or soil-like
form)
• to reduce environmental impacts from leached contaminants of drilling
wastes to aid transportation safety (i.e. to make loads more homogeneous for
transportation).

Solidification involves the addition of materials to the waste that physically


encapsulate or surround contaminants, such as cement and lime for oily drilling
wastes, decreasing the surface area across which pollutants can leach and
reducing the physical mobility of environmental contaminants. Solidification is
intended to improve waste handling and physical characteristics, while physically
limiting hazardous constituents (Barnes & Hartley, 2005), (Amiry, et al., 2008).

Stabilization is the conversion of a waste to a chemically stable form, based on the


reaction of reagents with waste and water to promote sorption, precipitation or
incorporation into crystal lattices. The technique is suitable to immobilize watery
sludges to yield a powdery hydrophobic product, which can be compacted, and may
be accomplished by pH adjustment to create a low leaching-risk product.

Stabilization or solidification can be used separately or in conjunction depending


on the required end product. Fundamental mechanistic differences between the
processes exist, but they are considered together as a remediation technique due
to shared outcomes afforded by each method.

Solidification/stabilization involves the addition of reagents to drilling waste,


requiring a mixing method, typically in situ backhoe mixing (which uses minimal
equipment and manpower), or ex situ ploughshare mixing, reported to increase
efficiency in the mixing process and maximize waste-to-additive contact (Scomi
Oiltools, 2008). A range of mixing plants exist from small scale plant using fully
automated systems, to lined waste pits mixed using a backhoe.

An example of a cuttings stabilization process is shown in Figure 34 selected steps


in a cuttings stabilization process are shown in Figures 35 – 37.
Drilling waste management technology review 63

Figure 34: Cuttings stabilization process (M-I SWACO)

Figure 35: Lined concrete pits Figure 36: Mixing/blending unit Figure 37: Treated and
to receive the untreated cuttings (Photo by Chevron) Stabilized Cuttings
(Photo by Chevron) (Photo by Chevron)

Automated solidification of liquid drilling wastes is demonstrated by one vendor in


which solidification occurs by:
• homogenizing waste drill cuttings and keeping solids in suspension in a roll-
on/off container
• pumping the slurry over a shaker to remove large cuttings
• mixing a pre-determined mixture of reagent in from a feed-hopper, and
• pumping treated discharge into a second (storage) roll on/off container,
where the reagent–waste mix starts binding the material, becoming solid
(TASC, 2015).
Drilling waste management technology review 64

The drilling fluid used affects options for disposal or use of cuttings, which retain
some drilling fluid (Freidheim & Candler, 2008). Academic research addresses
the efficacy of using the drill cutting solidified/stabilized products using differing
combinations of reagents (Al-Ansary & Al-Tabbaa, 2007) (Burnett, 2014) (Barnes
& Hartley, 2005) (Energy Pipeline, 2015). Additionally, vendors have reported that
drill cuttings arising from water-, oil-, or synthetic-based fluid operations can be
stabilized (Halliburton, 2007).

Portland cement is a common reagent and other additives such as power plant
fly ash, blast furnace stabilization/solidification slag, cement kiln dust, calcium
oxide, bentonite and other binding agents are widely used alone or in conjunction
with cement (Leonard & Stegemann, 2010a) (Ball, et al., 2011). Additionally, a study
has shown improved immobilization of organic contaminants where cuttings from
synthetic based fluid operations are mixed with high carbon fly ash as an additive
to Portland cement mixtures (Leonard & Stegemann, 2010).

Research has identified that differing ratios of dry binder-to-water can be used in
stabilization/solidification mixes. Drill cuttings mixes with varying oil content have
been researched with 0.4:1 to 0.6:1 water to binder ratio, with a greater content of
dry binder proving most effective in stabilizing drill cuttings from synthetic based
fluid operations (Al-Ansary & Al-Tabbaa, 2004) (Al-Ansary & Al-Tabbaa, 2007).

The use of stabilized drill cuttings for road construction and well pad material
has been demonstrated in Venezuela (Halliburton, 2007) and Texas, USA (Burnett,
2014), amongst other sites, with success. In Colorado, USA, stabilized cuttings
which met the Colorado Department of Transportation’s classification for road
base were produced from water-based drill cuttings (Energy Pipeline, 2015).
Studies have demonstrated the use of solidified/stabilized drill cuttings in road
base materials and spine roads, highlighting that sampling at the beginning of
the project and thirteen months into the project yielded similar concentrations of
metals (Burnett, 2014).

Alternative construction-related examples of solidified/stabilized drilling wastes


include levee construction, demonstrated in Louisiana (Gonzales, et al., 2008),
earthworks backfill, landfill lining materials and landfill daily cover (Leonard &
Stegemann, 2010a). Stabilization/solidification products created from Portland
cement, with the addition of high-carbon power plant fly ash (HCFA), have been
shown to be capable of achieving UK acceptance criteria for management as
non-hazardous waste, however chloride and oil content preclude products being
classified as inert (Leonard & Stegemann, 2010a). Another study showed the
potential application of solidified/stabilized waste as “Controlled Low Strength
Materials” (flowable fills) deemed self-compacting cement-based materials, used as
alternatives to conventional compacted fills (The American Concrete Institute, 1999).
Drilling waste management technology review 65

Solidification can allow disposal of bulk cuttings (formerly) containing liquid at


some licensed municipal solid waste landfill sites in the United States (Ohio
Environmental Protection Agency, 2012) and can assist in the safe transportation
of wastes by constraining the movement of water weight in waste-carrying vehicles
(Energy Pipeline, 2015), providing health and safety advantages.

Brick and building materials are noted as an attractive cuttings disposal route
when using Portland cement as a conventional binder (Ogechi Opete, et al., 2010)
(El-Mahllawy & Osman, 2010). Furthermore, leaching decreased with increasing
amounts of dry-binding material, allowing the drill cutting waste to be classified as
non-reactive and non-hazardous (Ogechi Opete, et al., 2010).

A study regarding stabilization of North Sea synthetic-based drill cuttings,


(containing hydrocarbons up to 4.2% w/w and chloride content up to 3.32% w/w)
showed that cuttings could be reduced to a stable non-reactive hazardous state,
when conventional (Portland cement, blast furnace slag) and novel (micro-silica,
magnesium oxide) binders were tested (Al-Ansary & Al-Tabbaa, 2007). Cuttings
were homogenized, mixed with water and varying quantities of dry-binder (10-
30%), and left to cure in molds for 28 days. Similar studies have used mechanically
separated drilling wastes, with stabilization/solidification proposed as an
alternative to other more energy intensive or expensive solutions such as landfill
disposal (Al-Ansary & Al-Tabbaa, 2004).

Contaminants contained in drill cuttings are dependent on the drilling fluid used
and the formation geology (Leonard & Stegemann, 2010a) and stabilization and/
or solidification relies on the correct mixture of immobilizing reagents (Al-Ansary
& Al-Tabbaa, 2007) (Ball, et al., 2011). Studies have highlighted that the correct
selection of reagents has a direct impact on the engineering properties of the end-
product (Amiry, et al., 2008) (El-Mahllawy & Osman, 2010) (Burnett, 2014); hence,
strength and suitable pH, metal, chloride, and hydrocarbon leaching testing needs
to be carried out to ensure product stability, with continued monitoring of products
for two years proposed in some cases (Barnes & Hartley, 2005).

Another process involves the application of a range of charcoal based products


which the vendor states can offer adsorption of leachable organic contaminants,
metals and metalloids to meet environmental quality standards, and retain these
contaminants across a broad range of redox and pH conditions. The vendor reports
that the process has been trialed successfully on drilling wastes at sites in Asia
(C-Cure, 2012).

Though there are many potential uses for recycled drill cuttings waste, long-term
leaching effects are noted in some matrices which have been shown to break down
in real environments (Khodja, et al., 2007). Cement fixation is considered to be best
suited for inorganic waste (Ifeadi, 2004), and the composition of the drill cuttings
Drilling waste management technology review 66

material can pose limitations on the applicability of stabilization/solidification in


cement-based systems, including when;
• the organics content is above 45% by weight
• the wastes have less than 15% solids
• excessive quantities of fine soil particles are present, and/or
• too many large particles are present (Flemming, 2000).

Difficulties with the use of cement or lime alone in the treatment of organic
contaminants have been reported as organic compounds can interfere with the
hydration of cement/lime, resulting in retardation of the hydration process and a
reduction in material strength.

Considerations of the embodied energy needed to manufacture the cement


and materials used are mentioned, noting that the volume of cement and costs
of binding materials needed in some instances outweigh the benefits of using
solidification/stabilization over other methods of drilling waste management, for
example, thermal desorption (Page, et al., 2003).

Following solidification/stabilization, drilling waste is potentially suitable for a


variety of beneficial construction purposes either onsite or offsite, and as such
the technology provides an avenue to reduce environmental effects of drilling
operations by transporting less material away from site and either diverting waste
from landfill or allowing waste to meet landfill acceptance criteria.

Cavitation scrubbing
Cavitation scrubbing works on the basis of subjecting a waste stream to rapid
changes of pressure. Vapour bubbles form in lower pressure regions, which on
collapse release significant amounts of energy and high-velocity micro-jets of
liquid which are reported to separate hydrocarbons and fine particulates from
larger solids. Information is provided by suppliers in a technical data sheet (Global
Advantech, 2012) but no published literature or case studies were identified which
describe this system in use.

Supercritical fluid extraction


Supercritical fluid extraction uses substances at or above their critical pressure
and temperature as solvents. In the vicinity of the critical point, the liquid and
vapor phases of the substance merge, producing a fluid with gas-like diffusivity
and viscosity and liquid-like density. These properties favor the transfer of soluble
waste components (i.e. hydrocarbons) from solid matrices to the supercritical
fluid. The density of the fluid is defined by the pressure and temperature; small
changes in processing conditions can fine-tune the solvating power of the fluid.
Drilling waste management technology review 67

Research into the use of supercritical carbon dioxide for the treatment of OBM
cuttings has been carried out by the University of Alberta (Street & Guigard,
2009). The process is carried out at temperatures which minimize the potential
for cracking of the recovered drilling fluid (40 – 50ºC in trials) and these trials
have reduced oil content from 17 – 19% to 0.3 – 0.6% (Street & Guigard, 2009). The
University of Alberta is currently collaborating with an oil field services company to
commercialize this technology.

Other processes
A variety of other processes have been trialed or are under development. These
include an Australian vendor which has developed a continuous flow process, which
uses a multi-stage physical processing system to dewater drill cuttings, fluids and
other liquid wastes. The vendor has carried out trials and states that the system
has potential applicability to a wide range of wastes, but the precise techniques
used in this system are not in the public domain (CPM Engineering, 2014).

Disposal

Offshore discharge
Offshore discharge of drill cuttings from offshore drilling operations may be
permissible provided the cuttings comply with the relevant discharge criteria (see
section 3 for a brief overview for selected jurisdictions).

When compared to transferring drilling waste to shore for treatment, offshore


discharge has the following potential benefits:
• reduced safety risk (avoiding movements of cuttings boxes)
• reduced environmental impacts associated with vessel and vehicle
movements
• high availability, since offshore discharge is less likely to be constrained by
adverse weather or vessel availability and hence has low risk of interruption
• reduced transportation costs.

For these reasons, offshore discharge is usually preferred where permitted, where
no major environmental impacts are expected, and where the necessary treatment
standards can be achieved cost-effectively.
Drilling waste management technology review 68

Re-injection
Cuttings reinjection (CRI) is a commonly used treatment method for both economic
and environmental reasons (Svenson & Taugbol, 2011).

Cuttings are slurrified at the surface to reduce particle size and produce an
injectable mixture (Figure 38).

Figure 38: Key component of the cuttings reinjection process (Chevron)

The cuttings slurry is then hydraulically injected into a subsurface formation that
is receptive and permanently isolated at a safe depth to prevent propagation to
ground water resources or to the surface (Alba Rodriguez, et al., 2007).

The advantages of CRI are:


• it is consistent with zero discharge regulations, with no direct releases to the
surface environment, and
• it reduces the logistical burden, with no need to ship cuttings to remote
disposal facilities.

CRI can also be a very cost-effective solution. It is particularly well suited to


remote locations where alternative waste disposal infrastructure is lacking, and
sensitive environments where zero discharge is required (Guo & Nagel, 2009).
Drilling waste management technology review 69

Slurrified cuttings can be injected either in a dedicated injection well, or via the
annulus of an existing production or water injection well (Gumarov, et al., 2014).

The CRI system includes the following components (Alba Rodriguez, et al., 2007):
• cuttings transport system
• slurrification system
• slurry storage tanks, and
• re-injection system.

Transportation systems include: gravity collection, augers/belts, vacuum systems


or pneumatic transfer. Although simple, the use of gravity collection is usually be
precluded due to space limitations. Auger and vacuum systems are effective but
can be limited by distance and layout. Pneumatic systems allow transport over
considerable distances and elevation changes, and as contained systems can
minimize the likelihood of spillage (Alba Rodriguez, et al., 2007).

Slurrification is carried out by first mixing the cuttings with water in a coarse
tank, and then circulating using centrifugal degradation pumps to degrade the
solids. The slurry is then pumped to a classification shaker and grinder where the
remaining particles are reduced in size by a grinder. The slurry is conditioned by
adjusting the solids to liquid ratio and adding any necessary chemicals, and then
transferred to a holding tank for reinjection. It is reported that normal practice
is to grind particles to a size of <300 µm, and to form a slurry with 20% or less of
solids by volume, although larger particle size and higher solids concentrations
may be feasible depending on the actual properties of both the slurry and receiving
formation (Ji, et al., 2010).

Cuttings are then re-injected into the chosen formation using a high pressure
injection pump, which creates fractures in the formation accepting the cuttings
slurry.

The use of cuttings reinjection as an alternative to skip-and-ship is reduces


the number of crane lifts required for cuttings boxes, offering health and safety
benefits; it also allows for zero discharge of cuttings to the surface environment.
(Reddoch, 2000).

CRI is potentially applicable for operations at onshore facilities or fixed offshore


platforms. Reports of a few trials of CRI from deep water drilling rigs have been
published. (Guo, et al., 2007). Mobile slurrification systems are also available
for onshore operations, which can be used to re-inject stockpiled cuttings into
redundant wells.
Drilling waste management technology review 70

The main challenges of CRI are (Guo, et al., 2007):


• the need for conducting geo-mechanical studies
• regulatory and permitting issues
• ensuring proper containment of injected cuttings
• rheology design of slurry to be injected
• injection and displacement procedures
• capacity of disposal well
• reliability of surface equipment (including grinders)
• monitoring and verification.

The response of the receiving formation (including fracturing characteristics and


containment zones) is simulated by modelling prior to injection.

The modelling process seeks to characterize the:


• containment barriers in the formation, based on stress, permeability or
modulus barriers which limit the extent of fracturing, and
• mechanisms for storage in the receiving formation, including how the
formation behaves when subjected to repeated slurry injections over time
(Guo, et al., 2007).

Important factors in slurry design are:


• viscosity of the slurry
• cuttings suspension, which must be sufficient to prevent cuttings settling out
of the slurry during the period it may remain in suspension in the well
• particle size, which must be sufficiently small to prevent plugging of the well,
or of the fractures close to the wellbore, either of which will limit further
injection (Guo, et al., 2007).

It is reported that the majority of complications during CRI are a result of wellbore
plugging (Gumarov, et al., 2014). Both corrosion and erosion can cause damage
in dedicated injection wells used for long-term injection projects (Gumarov,
et al., 2014). Erosion can be mitigated by selection and design of tubulars and
corrosion by the use of oxygen scavengers and displacement to suspension fluid if
injection stops for a prolonged period of time (Gumarov, et al., 2014). Good quality
cementing is also important to prevent loss of containment by channeling of fluids
behind the casing as a result of poor cement bond quality in dedicated injection
wells (Gumarov, et al., 2014).

CRI has the potential to deal with very large volumes of waste. For example, CRI
was selected as the preferred disposal method for 1.5 million bbl of drilling waste
expected to arise from the Karachaganak field in Khazakstan; and 3.5 million bbls
is reported to have been injected into a single well in the Ekofisk field in the North
Drilling waste management technology review 71

Sea (Nagel, 2005). Elsewhere, CRI is used in the US, China, North Sea, South
America, West Africa, Russia and Central Asia (Gumarov, et al., 2012). Injection
rates are reported to be typically in the range of 4 – 5 bbl/min, which is sufficient
for propagating hydraulic fractures, safely emplacing the waste, and avoiding
excessive erosion of the well and equipment (Gumarov, et al., 2012).

Potential risks in CRI fall into two categories: those related to subsurface
conditions; and those related to the surface operations of slurry processing and
pumping. Failures in either can stop the drilling operation, and hence it can be
beneficial to allow for cuttings storage within the operation so as to allow drilling
to continue even if there are temporary problems with slurrification or injection
(Guo & Nagel, 2009). Injectivity tests can be performed on the well to determine
the actual properties and validate modelled parameters. Injection pressure
monitoring and performance assessment during injection is then used to identify
any developing risks (Guo & Nagel, 2009).

The impacts of CRI failures can be significant. In 2009, problems were encountered
with several reinjection wells in the Norwegian Coastal Shelf, resulting in fractures
extending up to the seabed, some wells being closed, and others operating with
limitations on injection rates and volumes. This resulted in a considerable increase
in volumes needing to be shipped to shore (Svenson & Taugbol, 2011).

Onshore landfill
Disposal of waste by means of burial onshore is generally referred to as landfilling,
which is widely used throughout the world for the final disposal of a wide range of
wastes.

The main principle of landfills (as opposed to land application of waste) is to


ensure that wastes are contained (by the geology of the site and/or engineering
measures) such that the waste itself and the products of waste decomposition
do not have an adverse impact on the surrounding environment. A facility of this
type is commonly referred to as an ‘engineered landfill’ or ‘sanitary landfill’, to
distinguish them from sites where waste is dumped with no measures to protect
the environment. Different countries have different specific requirements for
landfill site selection, construction and management.

The European Union (EU) Directive on the Landfilling of Waste defines landfill as “a
waste disposal site for the deposit of the waste onto or into land (i.e. underground),
including internal waste disposal sites (i.e. landfill where a producer of waste is
carrying out its own waste disposal at the place of production), and a permanent
site (i.e. more than one year) which is used for temporary storage of waste”. The
EU definition of landfill, however, excludes “the use of inert waste which is suitable,
in redevelopment/restoration and filling-in work, or for construction purposes,
in landfills…[and] the deposit of unpolluted soil or of non-hazardous inert waste
Drilling waste management technology review 72

resulting from prospecting and extraction, treatment, and storage of mineral


resources as well as from the operation of quarries.” (European Union, 1999).

The EU Landfill Directive classifies landfills into one of the following categories:
• landfill for hazardous waste
• landfill for non-hazardous waste, and
• landfill for inert waste.

In the United States, non-hazardous waste landfills are regulated by the individual
states, but in compliance with Title 40 of the Code of Federal Regulations (CFR)
258: Criteria for Municipal Solid Waste Landfills (commonly referred to as ‘Subtitle
D’). While many E&P wastes are exempt from regulation as hazardous waste under
the Resource Conservation and Recovery Act (RCRA) Subtitle C, these wastes are
generally subject to non-hazardous waste regulation under RCRA Subtitle D and
applicable state regulations (USEPA, 2014).

The International Finance Corporation (IFC) publishes a series of Environmental,


Health and Safety (EHS) Guidelines, which describe the standards expected
of projects receiving IFC funding. IFC define ‘sanitary landfill’ as ‘a carefully
engineered, structurally stable formation of segregated waste cells separated by
soil cover material, with base and side slopes designed to minimize infiltration and
facilitate collection of leachate.’ (International Finance Corporation, 2007)

For general landfills (i.e. non-hazardous), IFC refers to applicable national


requirements and internationally recognized standards (including US EPA and EU
Landfill Directive).

For hazardous waste landfills, IFC recommends:


• a bottom lining system, preferably with two or more low-permeability liners;
• a leachate collection system above the upper liner to limit leachate depth to
0.3 m, and
• in the case of double lined systems, install leak detection between the two
liners.

US EPA effluent guidelines for leachate treatment are referenced as benchmarks.


IFC recommends that hazardous waste landfills are sited:
• no closer than 250 m from residential development
• no closer than 500 m from down-gradient water sources
• outside the 10 year flood plain, and
• avoid landfill development on outcropping limestone (or other carbonate
rock), within 500 m of a fault, or within the 10-year recharge area of water
supply.
Drilling waste management technology review 73

IFC also recommends installation of a landfill gas collection system in accordance


with applicable national requirements and internationally recognized standards.
This includes either recovery or flaring, with provision for condensate removal.
The use of blowers is recommended, which implies active rather than passive
venting. Hazardous waste landfills must be provided with a cap having equal or
lower permeability than the basal liner. Installation of groundwater monitoring
wells outside the landfill perimeter is also recommended (International Finance
Corporation, 2007).

Some jurisdictions have permitted drilling waste to be disposed of in situ at the rig
site, by burial of the drilling pit which contains the cuttings. The extent to which
this approach is permitted and the necessary conditions (e.g. use of impermeable
liners) will vary between jurisdictions. For example, in the United States, individual
States publish guidelines on acceptable drilling pit management methods
(Colorado Oil & Gas Conservation Commission, 2014) (Railroad Commission of
Texas, 2015).

The International Finance Corporation’s Environmental, Health and Safety


Guidance for onshore oil and gas development recommends the following
minimum conditions for in situ burial of drill cuttings in the drilling pit
(International Finance Corporation, 2007):
• The pit contents should be dried out as far as possible.
• If necessary, the waste should be mixed with an appropriate quantity of
subsoil (typically three parts of subsoil to one part of waste by volume).
• A minimum of one metre of clean subsoil should be placed over the mix.
• Topsoil should not be used [as the main component of the cap] but it should
be placed over the subsoil to fully reinstate the area.
• The pit waste should be analyzed and the maximum lifetime loads (i.e. total
amounts of chemicals added to the soil over the lifetime of the drilling
programme) should be calculated. A risk-based assessment might be
necessary to demonstrate that internationally recognized thresholds for
chemical exposure are not exceeded.

Landfill disposal is a relatively simple, low-cost technology for solid wastes.


However, the waste remains in situ within the landfill and there is the potential for
groundwater contamination if burial is not done correctly or contaminated wastes
are buried. The waste producer or landfill operator may therefore retain liabilities
for monitoring and potential remediation for an indefinite period.
Drilling waste management technology review 74

Salt cavern disposal


Salt caverns used for oil field waste disposal are created by a process called
solution mining. Well drilling equipment is used to drill a hole from the surface
to the depth of the salt formation and a smaller diameter pipe is lowered through
the middle of the well. This arrangement creates two pathways, into and out of
the well. To form a salt cavern, the well operator pumps fresh water through one
of the pipes. As the fresh water comes in contact with the salt formation, the salt
dissolves until the water becomes saturated with salt. Cavern space is created by
the removal of the salt-laden brine.

Wastes are brought to the cavern site in trucks and unloaded into mixing tanks
where they are blended with water or brine to make a slurry. Many exploration and
production wastes are suitable for disposal in caverns, including drilling fluids,
drill cuttings, produced sands, tank bottoms, contaminated soil, and completion
and stimulation wastes. Grinding equipment may be used to reduce particle size.
The waste slurry is then pumped into the caverns. The incoming waste displaces
the brine, which is brought to the surface and either sold or injected into a
disposal well. Inside the cavern, the solids, oils, and other liquids separate into
distinct layers: solids sink to the bottom, the oily and other hydrocarbons float to
the top, and brine and other watery fluids remain in the middle (Drilling Waste
Management Information System, undated).

A vendor serving customers in Western Canada offers underground salt caverns


as an environmentally sound disposal option for oilfield liquid and solid wastes
(Tervita, undated). Waste products are pumped into the caverns, typically in slurry
form, displacing the naturally occurring brine, which is removed and re-injected
into disposal wells. The vendor reports that they operate salt cavern disposal
facilities in Lindbergh, Alberta and Unity, Saskatchewan, and that the cavern
operations are licensed to accept a wide range of upstream oil and gas waste.

Salt cavern disposal is also used in United States, with a company operating in
Texas, Louisiana and New Mexico servicing the Eagle Ford Shale, the Permian
Basin, the Barnett Shale and the Haynesville Shale areas and interest has grown
for siting new commercial disposal caverns near the coast that can receive wastes
from offshore operations. Non-hazardous oilfield waste is pumped from unloading
facilities into the bottom of the salt cavern.

The operator states that the benefits of this disposal route are:
• salt caverns are located well below groundwater and separated from
freshwater sands by cap rock formation;
• salt domes are extremely stable geologically; and
• salt is virtually impermeable and behaves as a plastic rather than brittle
material (Trinity Environmental Services, 2015).
Drilling waste management technology review 75

Application to land
Land application in the context of this review is defined as directly applying drill
cuttings to land in a controlled manner, where the drill cuttings are intended to
be incorporated into the natural soil structure. This definition excludes landfilling
(where waste is placed in a designated area that is isolated from the surrounding
environment) and dumping (uncontrolled disposal of waste).

For the purposes of this review, composting (biodegradation of waste in a


controlled and isolated system) is considered separately, whereas biodegradation
of waste within a natural soil system is considered as a form of application to land.

The terminology used for the different forms of application to land varies between
different jurisdictions.

As an example, the classification of methods adopted in Alberta, Canada is as


follows (Alberta Energy Regulator, 2012):
• land-spray – spraying fluid or total waste onto topsoil and may or may not
involve incorporating the waste into the soil
• land-spray while drilling (LWD) – spraying drilling wastes onto topsoil at
controlled, low-application rates. LWD is limited to drilling wastes from
water-based drilling fluid systems, which are shown to be nontoxic using the
luminescent bacteria toxicity test
• pump-off – pumping the clear liquid portion of drilling wastes onto land
(usually vegetated land) using irrigation equipment
• mix-bury-cover – mixing drilling waste solids or total waste with sub-soils to
form a stabilized waste/soil mass that resides below the major rooting zone
• land-spreading: spreading drilling waste on the shallow subsoil and
incorporating it into the shallow subsoil
• Land Treatment – applying drilling waste to a dedicated parcel of land and
cultivating it into the receiving subsoil, where the inherent soil processes
biodegrade, transform, and assimilate the waste constituents – this is often
referred to as land-farming.

Application to land can prove very cost-effective since the equipment requirements
are relatively small. The main constraints are regulatory: the feasibility of
application to land varies greatly depending on the regulatory structure and
requirements of the country in question, sometimes with considerable variation
between jurisdictions within the same country.
Drilling waste management technology review 76

The three main determinants which influence the suitability of this technique and
the application rate are:
• salt content – increasing the salt content of land can be harmful to plant
growth and the productivity of agricultural land, as well as affecting
groundwater
• heavy metal content – although barium (from barite) has limited
bioavailability, more problematic heavy metals such as lead and mercury may
be present in the waste
• oil content – either entrained formation hydrocarbons or residual NADF.

Land-spray
The land-spray disposal method involves spraying fluid or total waste onto topsoil
and might or might not involve incorporating the waste into the soil. Incorporation
is typically done when the drilling waste has been land-sprayed on cultivated
land. It is accomplished by mechanically combining the drilling waste into a
homogeneous soil/waste mix. Drilling waste that has been land-sprayed on
vegetated land is typically not incorporated. A calculated loading rate or maximum
application rate is used to determine the area required for land-spraying.

Land-spray while drilling


The land-spray while drilling (LWD) disposal method involves spraying drilling
wastes onto topsoil at controlled, low-application rates. LWD is typically limited
to drilling wastes from water-based drilling fluid systems, which are shown to be
nontoxic using the luminescent bacteria toxicity test. The regulations in Alberta
also permit application of drilling waste from nontoxic, water-based fluid systems
onto forested lands, which may be used particularly during winter in areas of harsh
climate.

The land-spray and LWD disposal requirements are similar, but as the LWD
method is limited to only nontoxic fluid systems, the testing requirements have
been reduced, and disposal is allowed to proceed without first storing the drilling
waste (e.g., by sumpless drilling). The controlled application is normally conducted
during the drilling operation. Spraying techniques may include the use of vacuum
trucks or similar equipment.

Pump-off
The pump-off disposal method involves pumping the clear liquid portion of drilling
wastes onto land (usually vegetated land) using irrigation equipment, such as a big
gun, sprinkler, gated pipe, or perforated hose, or using water/vacuum trucks.
Drilling waste management technology review 77

Mix-bury-cover
The mix-bury-cover (MBC) disposal method involves mixing non-hydrocarbon-
based drilling waste solids or total waste with sub-soils to form a stabilized waste/
soil mass that resides below the major rooting zone.

Typical MBC methods are:


• to mix waste and subsoil in the sump and cover
• to mix waste from the sump and subsoil on the surface, then put the mixture
back in the sump and cover
• to bail the waste from the sump onto the lease surface, mix with the subsoil,
and bury when filling in a cut, or
• to spread the waste on the lease surface and allow it to dry, put the waste
back into the sump, then mix and cover.

Land-spreading
The land-spreading method involves spreading water-based drilling waste on the
shallow subsoil and incorporating it into the shallow subsoil.

Typical land-spreading methods are:


• to rip the subsoil and spread (squeeze) and incorporate the waste from the
sump on site
• to spread (squeeze) the waste from the sump on site, dry it, and then
incorporate it into the shallow subsoil.

Land Treatment/Land-farming
Land Treatment involves applying the drilling waste to a dedicated parcel of land
and cultivating it into the receiving subsoil, where the inherent soil processes
biodegrade, transform, and assimilate the waste constituents. As a technique,
application to land may be particularly applicable to remote on-shore locations,
where alternative management techniques are limited and where there is plenty of
available land which can be used for application, and the rates of application can
be carefully controlled and monitored.

Drilling waste managed by land treatment becomes assimilated within the soil in
which it was mixed. To maximize the success of land treatment and to minimize
soil and environmental impacts, the application rate and incorporation depth of the
drilling waste are limited. The method requires tillage and application of nutrients
to break down the petroleum hydrocarbon in the drilling waste. Small amounts of
amendments (e.g. nutrient, organic) may be added to enhance the biodegradation
process. Sampling and analyses are necessary to monitor the progress of
biodegradation. This process is very similar to composting except that it is carried
out in natural soils, rather than in a contained windrow, pile or vessel.
Drilling waste management technology review 78

Management of salt contaminated cuttings


Drill cuttings can be contaminated with salt both as a result of using salts in
drilling fluids, and also as a result of drilling through formations containing salts.

The problems associated with salt in cuttings are:


• any breach in containment of pit liners (when using drilling pits) can
contaminate groundwater
• salt is not biodegradable and hence salt content is not reduced over time by
microbial action in the same way as hydrocarbon content
• salt is readily soluble and hence highly mobile in the environment, and
• high salt concentrations can be highly toxic to plants and soil organisms, and
deleterious to soil structure (Clements, et al., 2008).

Several approaches are available for removing salts from drill cuttings:
• mechanical washing, using freshwater to wash out the salt content as part of
the solids control process. This, however, results in a stream of saline water
which then requires treatment itself
• leaching pad, where cuttings are placed on a lined leach pad and salts are
leached out with freshwater. The resulting salty water can be re-injected or
incorporated back into the fluid system, and/or
• addition of calcium ions – it is reported that adding calcium acetate or
gypsum to wash solution can enhance sodium chloride recovery during
washing or leaching by promoting ion exchange between calcium and sodium
ions (Clements, et al., 2008).

Beneficial reuse

Construction materials
Treated drill cuttings have been proposed or used for a variety of construction-
related purposes, including as:
• fill material
• road material
• daily cover material at landfills
• aggregate or filler in concrete, brick, or block manufacturing.

This application is closely linked to the techniques referred to in previous sections for
solidification and stabilization, as the cuttings often require stabilizing prior to use.
Drilling waste management technology review 79

In addition to the aforementioned applications, a research project was carried out


into the use of NADF cuttings as an aggregate replacement in cold mix asphalt,
as used in road pavement construction (Allen, et al., 2007). Constraints included
the high moisture content and hydrocarbon residues of the cuttings, which were
overcome by redesigning the type of binder used in the mix, and replacing a
proportion of the virgin aggregate with incinerator bottom ash. Trials were carried
out at a quarry, and the results were deemed to indicate the potential applicability
of this process. However, the variability of cuttings was noted as a potential
constraint. Other research has compared a range of samples of cuttings from
the North Sea against existing standard requirements for fillers in bituminous
mixtures. The materials were either readily suitable for inclusion in pavement
asphalt or after minor adjustment of the processing technology (Dhir, et al., 2010).

Research has also been carried out on the vitrification of drill cuttings, using a
process involving mixing with sodium and calcium oxides and heating at 1300ºC
for 5 hours, followed by further treatment at 750 – 800ºC. The resulting glass
ceramics were shown to exhibit high hardness and fracture strength and almost
zero porosity, making them potentially suitable for reuse in building applications,
e.g. as tiling (Abbe, et al., 2009).

A study in Taiwan assessed the potential for using drill cuttings in the manufacture
of building bricks and as a partial cement substitute in concrete. The researchers
reported that the resulting bricks were of sufficient compressive strength and
permeability for building purposes, and the concrete showed variable results and
was sensitive to the substitution ratio (Chen, et al., 2007).

Fuel
The use of drilling waste as a feedstock for cement kilns is mentioned frequently in
the general literature (National Petroleum Council, 2011) although there is limited
information available on practical implementation. Some oil and gas companies
refer to the use of significant quantities of hazardous waste in cement kilns
(Petrobras, 2012a).

There is reference in the literature to the use of drill cuttings as fuel, referring
to trials conducted in the United Kingdom using oily cuttings as a fuel at a power
plant. Cuttings were reported to be blended in at a low rate with coal, the primary
fuel source. The resulting ash was much the same as the ash from burning just
the coal (Drilling Waste Management Information System, undated).

There is no subsequent information in the literature to suggest that this approach


has been commercialized. Given the increasing adoption of more costly NADF as
alternatives to diesel and the ability to reduce BFROC to relatively low levels using
solids control equipment and dryers, it is doubtful whether the use of drill cuttings
as a fuel will be commercially adopted on a significant scale.
Drilling waste management technology review 80

Wetland restoration
Studies have been carried out into the potential use of drill cuttings as a source
of sediment for wetland restoration in Louisiana (Willis, et al., 2005), with positive
results. Despite the promising results of these trials, there is no indication in the
literature that they have been successfully implemented in the US or elsewhere.

Summary
The following table (Table 5) summarizes the applicability of the main technologies
and techniques for managing drill cuttings, and Figure 39 illustrates the indicative
range of BFROC achievable by the main forms of treatment (although it should be
noted that actual performance is very dependent on cuttings and fluid properties
may in some cases lie outside the ranges indicated).

Table 5: Summary of drill cuttings management techniques

Technique Portability Relative Commercialization Qualitative Advantages Constraints


cost environmental
impacts
Solids control
Shale Onshore/ $ Commercially Emissions and Robust BFROC of
shaker offshore proven environmental and proven resulting cuttings
impacts are low technology which is typically too
is a standard part high to allow for
of solids control direct discharge to
the environment
Vacuum Onshore/ $$ Commercially Emissions and Able to reduce Unable to achieve
shaker offshore proven environmental BFROC of NADF BFROC <1% hence
impacts are low cuttings to a level unable to meet
much lower than OSPAR discharge
a conventional criteria
shaker
Dryers
Cuttings Onshore/ $$ Commercially Emissions and Robust Generally unable
dryers offshore proven environmental and proven to achieve BFROC
impacts from technology <1% hence unable
cuttings dryers capable of to meet OSPAR
are low: they significantly discharge criteria.
are electrically reducing BFROC Able to meet
powered and do of cuttings and discharge criteria
not give rise to recovering fluid in the US and
significant direct other countries
emissions to air or
water
Drilling waste management technology review 81

Technique Portability Relative Commercialization Qualitative Advantages Constraints


cost environmental
impacts
Vacuum Onshore/ $$ Commercially Similar to Capable of Generally unable
dryers offshore proven conventional reducing BFROC to achieve BFROC
cuttings dryers, to a lower level <1% hence unable
although slightly than standard to meet OSPAR
higher power dryers discharge criteria.
consumption Able to meet
discharge criteria
in the US and
other countries
Thermal treatment
Indirect Onshore $$$ Commercially Source of fuel Proven Unsuitable for
thermal only proven required to raise technology offshore use,
desorption heat, and requires capable of usually requires
good emissions achieving very fixed onshore
control to prevent low BFROC (<1%) plant
fugitive emissions
of volatile organics
and particulates
to air
Thermo- Onshore/ $$$ Commercially Fuel or Electrical Suitable for Heavy and space
mechanical offshore proven power required offshore use with consuming setup
desorption to generate heat, potential to meet for offshore
and requires good OSPAR discharge installation,
emissions control criteria (BFROC requiring
to prevent fugitive <1%) engineering
emissions of studies to retrofit
volatile organics on existing rigs
and particulates
to air
Microwave Not Under No information Potentially Not commercially
thermal proven development by available on suitable for proven
desorption oil field services environmental offshore use,
companies performance – will but yet to be
require electrical commercially
power to generate proven
microwaves
Incineration Onshore $$$ Commercially Requires good Proven Fluid is
only proven emissions control technology combusted
to prevent fugitive capable of almost and hence not
emissions of complete removal recoverable.
volatile organics of BFROC Costly air pollution
and particulates control equipment
to air may be required
Drilling waste management technology review 82

Technique Portability Relative Commercialization Qualitative Advantages Constraints


cost environmental
impacts
Biological treatment
Windrow/ Onshore $ Commercially Energy Simple and Addition of
static pile only proven requirements low-cost, able to organic material
composting are low. Effective achieve very low will increase
environmental endpoints overall volume of
management waste. Relatively
is required to large space
prevent fugitive requirements,
releases, and applicability
especially to limited in very
surface and cold climates.
groundwater Treatment
process may take
considerable
time (weeks
or months to
achieve required
endpoints)
In-vessel Onshore $$ Commercially Similar to Simple and low- Applicability
composting only proven windrow/pile cost (although limited in very
composting, more costly than cold climates.
but use of windrow/static Treatment
containment pile), able to process may take
vessel reduces achieve very low considerable
risk of endpoints time (weeks
environmental or months to
releases achieve required
endpoints)
Physical and chemical treatment
Supercritical Not – Under – – Not commercially
fluid proven development by proven
extraction oil field services
companies
Solidification Onshore $$ Commercially Potentially allows Simple and low Long-term
and only proven for beneficial cost. Able to performance of
stabilization reuse and hence produce material material in the
moves waste up suitable for environment
the hierarchy. beneficial reuse needs careful
Care required evaluation to
to ensure long- minimize future
term stability of liabilities
materials in the
environment Larger volumes
generated
Drilling waste management technology review 83

Technique Portability Relative Commercialization Qualitative Advantages Constraints


cost environmental
impacts
Disposal
Offshore Offshore $ Commercially Impacts can be Lowest cost Applicability
discharge proven determined by option, where depends
toxicological permitted on cuttings
evaluation, but characteristics
applicability and regulatory
severely regime. Pre-
constrained in treatment may be
some jurisdictions required
for environmental
reasons
Re-injection Onshore/ $$ Commercially Potentially Low transport May require
offshore proven low, provided costs if re- drilling of
receiving strata injection well is dedicated
are effectively close to source reinjection well;
isolated from the of waste; high any blockage or
wider environment capacity and unplanned release
– in some cases, avoids discharge can be costly and
releases have to the wider time-consuming
occurred environment to resolve
Onshore Onshore $ Commercially Properly Established Regulatory
landfill only proven engineered technique which constraints
and managed can make use over type and
sanitary landfills of existing local quantities of
ensure isolation facilities material that
of contaminants can be landfilled;
from the wider high landfill
environment, taxes in some
but long-term jurisdictions.
environmental Potential long-
liabilities may term liabilities
remain, and if waste
may be more containment is
significant in the inadequate
case of burial pits.
Potential releases
to groundwater
may be of
particular concern
Salt cavern Onshore $ Commercially Few impacts High capacity and Limited to areas
disposal only proven provided integrity avoids discharge having suitable
of the cavern to the wider geological
is maintained, environment formations
but long-term
environmental
liabilities remain
Drilling waste management technology review 84

Technique Portability Relative Commercialization Qualitative Advantages Constraints


cost environmental
impacts
Application Onshore $ Commercially Careful Simple and low Regulatory limits
to land only proven management cost on BFROC, salt
and monitoring and metals
required to contents. Amount
prevent build- of waste applied to
up of non- any given amount
biodegradable salt of land is limited;
and heavy metals, hence large areas
and to ensure may be required
that hydrocarbons
are effectively
biodegraded
Beneficial reuse
Construction – na Commercially Potential benefits Environmentally May require
materials proven in substituting for benefits from considerable
virgin materials recycling waste, pre-treatment to
and reducing and reduces meet applicable
need for disposal/ costs of disposal specifications;
further treatment. requires approval
Need to ensure by regulators and
that materials end-users which
do not leach may be costly and
contaminants time-consuming
during use
Wetland – – No known practical Potential Potential Requires
restoration applications environmental benefits from regulatory
benefits in reduced costs approval –
facilitating and improved reported trials
ecosystem recycling
restoration
Cement kiln – – Commercially Potential Uses existing Fluid is
feedstock proven but limited environmental infrastructure combusted
information benefits in and no long-term and hence not
available replacing virgin liabilities if all recoverable
feedstock and fuel contaminants
are destroyed
and waste
transformed into
cement
Power – – No known current Not determined Not determined Fluid is
station fuel applications combusted
and hence not
recoverable
Drilling waste management technology review 85

BFROC content (%) 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20


Solids control

Shale Shaker

Vacuum-assisted shaker

Dryers

Vertical cuttings dryers

Vacuum dryers

Thermal treatment

Indirect thermal desorption

Thermo-mechanical
desorption

Incineration

Biological treatment
Windrow/static pile
composting

In-vessel composting

Physical and chemical treatment


Solidification and
n/a
Stabilization

Figure 39: Indicative BFROC performance of cuttings treatment techniques


Drilling waste management technology review 86

6. Management routes –
Interfacial mixtures and slops

This section describes the main technologies and methods used for managing
slops and interfacial mixtures.

General
‘Slops’ is a general term describing liquid mixtures of oil and water. Slops are
formed when drilling or displacement fluids, wash water from routine cleaning
operations, or rain water runoff become contaminated with drilling fluid
components. The slops are captured, either in surface pits or through the rig drain
system and collected in storage tanks (Mueller, et al., 2013).

The composition of slop can vary, but is reported to be generally in the range of
50–90% water and 10–50% oil (Ivan & Dixit, 2006), and may also contain up to 10%
solids (Mueller, et al., 2013).

Steps taken to reduce the volume of slops, including good housekeeping practices
(Mueller, et al., 2013) and technology to reduce mixing during fluid displacement
(Herigstad, et al., 2010) are consistent with the principles of good waste
management and can lead to reduced treatment and disposal costs.

The technologies used for treating slops are developed from those used for
treating other industrial wastewaters, often optimized to make them more rugged
and portable to meet the operational challenges of the drilling environment.

Slops treatment typically involves a two or more stage process. The degree of
treatment will depend on the discharge criteria for the treated water: for example,
the recommended limit on oil content set by OSPAR for discharges in the North
Sea is 40 parts per million (i.e. 40 mg/L) (Mueller, et al., 2013). In Brazil, the limit
on oil content for wastewater discharges is 20 mg/L (Cupelo, et al., 2014).

Various companies offer packaged slops treatment plants which are suitable
for use offshore or at onshore rig sites. There are also a number of land-based
treatment centers which are located to service established oil and gas operations.

Technologies used for treatment of slops can be split into two main types:
• chemical (e.g. flocculation)
• physical (e.g. dissolved air flotation and filtration).

Biological treatment appears to be less widely used as a primary means of slops


treatment, although may be applicable as part of a final polishing stage in a
wastewater treatment plant. Packaged biological treatment plants are available
Drilling waste management technology review 87

on the market (e.g. using membrane bioreactors), but are likely to be less robust
than physical/chemical treatment systems, particularly for wastewater where the
total organic content is relatively low – they are more typically used for treatment
of sewage or wastewater from food processing.

Many systems include multiple stages, often an initial chemical treatment to


break up the oil-water emulsion into distinct aqueous and non-aqueous phases,
followed by physical treatment to separate the two phases and further reduce the
oil content of the aqueous phase, to render it suitable for discharge. The choice of
treatment system is very dependent on the composition of the slop, and laboratory
or field tests are often needed in order to confirm the performance of a system
on a particular type of slop (Mueller, et al., 2013). Several companies supply
modularized units which are suitable for offshore use.

Chemical treatment
The two main objectives of chemical treatment are to:
• de-emulsify the oil/water mixture into separate oil and water phases, and
• cause the small colloidal (<2 micron) particulates to clump together
(flocculate) and hence allow for separation.

The types of emulsifiers and flocculants used will depend on a number of factors,
including whether the emulsion comprises water as the dispersed phase and oil
as the continuous phase, or the reverse (Mueller, et al., 2013). The presence of
fine colloidal particles can make the emulsion particularly hard to break. Oil field
services and specialty chemicals companies offer a wide range of proprietary
chemicals for this purpose.

The most common classes of water-in-oil demulsifiers are (Kelland, 2014):


• polyalkoxylate block co-polymers and ester derivatives
• alkylphenol-aldehyde resin alkoxylates
• polyalkoxylates of polyols or glycidyl ethers
• polyamine polyalkoxylates and related cationic polymers
• polyurethanes (carbamates) and polyalkoxylate derivatives
• hyperbranched polymers
• vinyl polymers
• polysilicones.
Drilling waste management technology review 88

Flocculants used in the oil industry comprise (Kelland, 2014):


• highly valent metal salts
• cationic polymers
• dithiocarbamates
• anionic polymers
• non-ionic polymers
• amphoteric polymers.

Following demulsification and flocculation, there are a number of techniques


which can be used to separate the solid, water and oil phases.

Gravity separation
Gravity separation relies on differences in density, viscosity and particle size to
separate the different phases, and is the principle behind the common ‘API separator’
that is widely used for treating oily water throughout the industry. Separation
efficiency can be improved by using baffles and plates within the separator. A
coalescing separator enhances separation by use of a coalescing medium, a
porous matrix such as fiberglass, mesh or wire wool (Mueller, et al., 2013).

The efficiency of gravity separation can be further enhanced by a number of


means, as presented below.

Dissolved air flotation


Dissolved air flotation (DAF) removes oil and solids from water by dissolving air
in the water under pressure and then releasing the air at atmospheric pressure
in a flotation tank or basin. The air bubbles which form adhere to the suspended
matter which causes it to float to the surface where it can be removed by
skimming (Ivan & Dixit, 2006).

Decanting centrifuge
A decanting centrifuge can perform either two-phase (solid/liquid) or three phase
(solid/oil/water) separation of the slop. In a two-phase centrifuge, the solids are
discharged at the ‘beach’ end of the centrifuge, with the liquid phase discharged
at the other end. In a three-phase centrifuge, the immiscible oil and water phases
are separated by density at the liquid end into two different discharge streams
(GEA Group, undated).
Drilling waste management technology review 89

Disc separator
A disc separator (or disc stack centrifuge) comprises a rotating bowl with special
plates (the ‘disc stack’), providing additional surface settling area, which contributes
to speeding up the separation process. The particular configuration, shape and design
of these plates make it possible for a disc stack separator to undertake the continuous
separation of a wide range of different solids from either one or two liquid phases.

Whereas decanter centrifuges are generally used for greater solids concentrations
with larger particle sizes, disc stack separators are used for separating lower
solids concentrations (e.g. <10% solids) and smaller particle and droplet sizes
(e.g. <100 microns) (Alfa Laval, undated).

Filtration
Filtration involves separating solids from liquids by passing through a medium
which retains the solid fraction. The effectiveness of the filter depends on the pore
size of the filter and the size of the particles. Several types of filter can be used in
slops treatment. Oil-absorbing filters used modified cellulose or organo-clay to
absorb oil; filter presses use mechanical pressure and filter plates; and activated
carbon filters adsorb contaminants (MI-SWACO, 2013).

Membrane separation
Membrane separation works on the basis of providing a physical barrier (in the
form of a membrane) that permits the passage of materials only up to a certain
size, shape or character.

Membrane separation technologies are generally categorized as ultrafiltration


(UF), reverse osmosis (RO), nanofiltration (NF), and microfiltration (MF).
• Ultrafiltration (UF) is a pressure-driven process that removes emulsified oils,
metal hydroxides, colloids, emulsions, dispersed material, suspended solids,
and other large molecular weight materials from water.
• The membrane with the smallest pores are used for reverse osmosis (RO),
which involves reversal of the osmotic process, using ionic diffusion to effect
the separation.
• Nanofiltration (NF) functions similarly to RO, but is generally targeted to
remove only divalent and larger ions.
• Microfiltration (MF) has the largest pore size and can be used to remove
bacteria and suspended solids.

Membranes are manufactured in a variety of configurations including hollow fiber,


spiral, and tubular shapes (Koch Membrane Systems, 2013).
Drilling waste management technology review 90

Ozone treatment
In a trial reported in the literature, a synthetic-based drilling fluid contaminated
with water was separated with a biodegradable demulsifier, resulting in separation
of the mixture into drilling fluid and water phases, but the separated water phase
was found to have an oil-in-water content of 4,000 mg/L which is unacceptable
for discharge. The separated water was then treated with ozone, and the ozone-
treated water was allowed to separate in two phases, i.e., clarified water phase
and organic rich phase. The oil-in-water content of the clarified ozone-treated
water phase was found to be 30 mg/L, thus reducing the oil-in-water content by
99% (Dixit & Patel, 2010).

Microwave treatment
Research has been carried out into the use of microwaves for demulsification of
oily sludges, but its application at an industrial scale is very limited (Hu, et al.,
2013). As of the time of reporting, there are not believed to be any commercial
applications of microwaves for slops treatment.

Electrocoagulation
Electrocoagulation is a process utilizing sacrificial anodes to form active
coagulants which are used to remove pollutants by precipitation and flotation in
situ. When contaminated water passes through the electrocoagulation cells, the
anodic process releases positively charged ions, which bind onto the negatively
charged colloidal particles in water resulting in coagulation. At the same time, gas
bubbles, produced at the cathode, attach to the coagulated matter causing it to
float to the surface where it is removed by a surface skimmer. Heavier coagulants
sink to the bottom, leaving clear water, suitable for use in drilling and production
operations.

Electrocoagulation is used commercially for treatment of produced and flowback


water and also has potential for treatment of slops.
Drilling waste management technology review 91

7. Management routes –
Other wastes

This section briefly describes the technologies and methods used for
managing other types of drilling wastes.

Drilling fluids

NADF
NADFs, because of their specialist nature and high cost, are typically returned to
the supplier once they are no longer usable for reconditioning and reuse.

WBDF
For WBDFs, various methods are available for managing fluids that are no longer
reusable. These include:
• direct discharge to sea
• application to land
• re-injection
• on-site dewatering
• treatment in an offsite wastewater treatment facility.

In the case of discharge to sea, re-injection, and application to land, similar


regulatory constraints apply as for drill cuttings. Where direct discharge of fluids is
not permitted or desirable, the operator must either treat fluids on site such that
part of the fluid can be discharged, or arrange for treatment at an offsite facility.

In the case of dewatering, the objective is to remove colloidal particles, producing


a ‘clean’ water phase that can be reused or that can meet regulatory requirements
for discharge. Most dewatering applications are reported to be required in zero
discharge regimes, where closed-loop systems are required. In closed-loop
systems, dewatering is the final step of the fluid-processing process, following
the separation of liquid from solids provided by the shale shakers, hydrocyclones
and centrifuges. Dewatering might also be necessary when there is a high risk
associated with fluid disposal, or where the volume of the active fluid system
needs to be reduced.

In a dewatering project, the drilling fluid will be processed; the resulting solids can
be managed in a similar manner to cuttings, with the clean liquid reused on site or
disposed to ground, if allowed.
Drilling waste management technology review 92

A fluid dewatering system will typically include a flocculation stage and a


separation stage (e.g. using a centrifuge). The addition of flocculants and
coagulants facilitate solids removal by the centrifuge (NOV, 2014).

Research has been carried out into the use of forward osmosis (FO) for the
treatment of aqueous oil and gas wastes, including drilling fluids. In forward
osmosis, a synthetic polymeric membrane separates a concentrated ‘draw
solution’ from a feed stream. The osmotic pressure difference across the
membrane facilitates diffusion of water across the membrane from the feed to
the draw solutions, with dissolved and suspended constituents being rejected.
The draw solution can then be subject to further treatment (including distillation
or reverse osmosis) to produce a clean water stream and a reconstituted draw
solution; the volume of feed solution requiring disposal is significantly reduced.

The claimed advantages of FO are the low energy requirements and capital costs,
and reduced membrane fouling when compared to pressure-driven membrane
systems such as ultrafiltration and reverse osmosis. A development of this
system, referred to as a ‘membrane brine concentrator’, is reported to have been
commercially demonstrated in the Marcellus Shale and Permian Basin in the
United States (Coday, et al., 2014).

WBFs can also be disposed of by mixing with a cross-linkable polymer and a


crosslinking agent to form a composition that solidifies either immediately or at
a predetermined time (Fink, 2015). Various other solidification agents such as
bentonite-based material and super absorbent polymer-based material can also
be used (National Driller, 2015).
Drilling waste management technology review 93

References

Aase, B. et al., 2013. Criticality Testing of Drilling Fluid Solids Control Equipment.
SPE Drilling and Completion, 28(02), pp. 148–157.
Abbe, O. E., Grimes, S. M., Fowler, G. D. & Boccaccini, A. R., 2009. Novel sintered glass-
ceramics from vitrified oil well drill cuttings. Journal of Material Science, 44(16),
pp. 4296–4302.
Al-Ansary, M. S. & Al-Tabbaa, A., 2004. Stabilisation/solidification of Synthetic North Sea Drill
Cuttings Containing Oil and Chloride. Barcelona, Spain, International RILEM Conference
on the use of Recycled Materials in Buildings and Structures.
Al-Ansary, M. S. & Al-Tabbaa, A., 2007. Stabilisation/solidification of synthetic petroleum
drill cuttings. Journal of Hazardous Materials, pp. 410–421.
Alba Rodriguez, A., Fragachan, F. E., Ovalle, A. & Shokanov, T. A., 2007. Environmentally
Safe Waste Disposal: The Integration of Cuttings Collection, Transport and Reinjection.
Veracruz, Society of Petroleum Engineers.
Alberta Energy Regulator, 2012. Directive 050: Drilling Waste Management, Calgary: Alberta
Energy Regulator.
Alfa Laval, undated. Alfa Laval – disc stack separator technology. [Online]
Available at: http://www.alfalaval.com/industries/food-dairy-beverages/food/
Documents/PCHS00022 Disc Stack Centrifuges.pdf
[Accessed March 2015].
Allen, B. et al., 2007. The development and trial use of oil exploration drill-cutting waste as an
aggregate replacement in cold-mix asphalt. Galveston, SPE E&P Environment and Safety
Conference.
Amiry, H., Sutherland, H., Martin, E. & Goodsell, P., 2008. Environmental Management and
Technology in Oil Refineries. In: S. T. O. (Ed.), ed. Environmental Technology in the Oil
Industry. 2nd ed. Hampshire, UK: Springer, pp. 307–308.
Aquateam COWI, 2014. Characterising Thermal Treated OBM Drill Cuttings, Oslo: Norwegian
Oil and Gas Association.
ASME Shale Shaker Committee, 2005. Drilling Fluids Processing Handbook. s.l.:Elsevier.
Ball, A. S., Stewart, R. J. & Schliepake, K., 2011. A review of the current options for the
treatment and safe disposal of drill cuttings. Waste Management & Research, 30(5),
pp. 457–473.
Barnes, G. R. & Hartley, D., 2005. Onsite Treatment of Oily Drilling Waste in Remote Areas.
Texas, Mud-loc USA Inc. American Association of Drilling Engineers.
Burnett, D. B., 2014. Low impact testing of oil field access roads: Reducing the environmental
footprint in desert ecosystems, Texas, US: Global Petroleum Research Institute.
Canadian Association of Petroleum Producers, 2001. Offshore Drilling Waste Management
Review.
Cannon, R. W. & Martin, D., 2001. Reduction of Synthetic Based Fluid Discharges Offshore
by the use of Vertical Basket Centrifuges. San Antonio, SPE/EPA/DOE Exploration and
Production Environmental Conference.
Drilling waste management technology review 94

C-Cure, 2012. Drilling Waste Treatment. [Online]


Available at: http://ccuresolutions.com/wp-content/uploads/2012/07/Drilling-waste-
TDS-Final1.pdf
[Accessed June 2015].
Chen, T., Lin, S. & Lin, Z., 2007. An innovative utilisation of drilling wastes as building
materials. Galveston, SPE E&P Environment and Safety Conference.
Clements, K., Fout, G., Marchbanks, W. & Rogers, D., 2008. Greenhouse-scale lab techniques
used to optimise removeal of chlorides from drill cuttings and qualify the feasibility of
beneficial reuse in the field. Houston, AADE Fluids Conference and Exibition.
Coday, B. D. et al., 2014. The sweet spot of forward osmosis: Treatment of produced water,
drilling wastewater, and other complex and difficult liquid streams. Desalination,
333(1), pp. 23–35.
Colorado Oil & Gas Conservation Commission, 2014. Policy on Drill Cuttings Management.
CPM Engineering, 2014. Ali-Jak Waste Treatment Process. [Online]
Available at: http://www.cpmengineering.com.au/_blog/CPM_Engineering_News/post/
drill-mud-waste-treatment/
[Accessed June 2015].
Cubility, 2014. This is the MudCube. [Online]
Available at: http://cubility.com/mudcube
[Accessed May 2015].
Cupelo, A. C. et al., 2014. Brazilian Deployment of Mobile Effluent Treatment of SBM
Contaminated Deck Drain and Slop Water Interfaces From Offshore Drilling Operations.
Long Beach, California, USA, SPE International Conference on Health, Safety, and
Environment.
DEFRA, 2007. Incineration of Municipal Waste, Waste Management Technology Brief.
Department of Food, Environment and Rural Affairs.
Dhir, R. K., Csetenyi, L. J., Dyer, T. D. & Smith, G. W., 2010. Cleaned oil-drill cuttings for use
as filler in bituminous mixtures. Construction and Building Materials, 24(3), pp. 322–325.
Dixit, R. & Patel, A. D., 2010. Design and Development of a Novel Process To Treat
Drilling-Fluid Slops: A Positive Environmental and Economic Impact. SPE Drilling and
Completion, 25(01), pp. 53–57.
Drilling Contractor, 2007. Solids Control: Shale shaker makers push advances in power,
efficiency, automation. [Online]
Available at: http://www.drillingcontractor.org/solids-control-shale-shaker-makers-
push-advances-in-power-efficiency-automation-1351
[Accessed May 2015].
Drilling Contractor, 2014. Solids control seeks balance of size, capacity. [Online]
Available at: http://www.drillingcontractor.org/solids-control-seeks-balance-of-size-
capacity-30679
[Accessed June 2015].
Drilling waste management technology review 95

Drilling Waste Management Information System, undated. Fact Sheet - Beneficial Reuse of
Drilling Wastes. [Online]
Available at: http://web.ead.anl.gov/dwm/techdesc/reuse/
[Accessed June 2015].
Drilling Waste Management Information System, undated. Fact Sheet - Disposal in Salt
Caverns. [Online]
Available at: http://web.ead.anl.gov/dwm/techdesc/salt/
[Accessed June 2015].
El Dhorry, K. & Dufilho, B., 2012. Automation Improves Shaker Performance. San Diego,
IADC/SPE Drilling Conference and Exhibition.
El-Mahllawy, M. S. & Osman, T. A., 2010. Influence of Oil Well Drilling Waste on the
Engineering Characteristics of Clay Bricks. Journal of American Science, 6(7),
pp. 48–53.
Energy Pipeline, 2015. Making Black Gold Fit for the road. [Online]
Available at: http://www.tracyhume.com/wp-content/uploads/2014/02/EnergyPipeline_
Feb15_RecycledDrillCuttings.pdf
[Accessed 28 April 2015].
Environment Agency, n.d. The Treatment of Waste by Thermal Desorption - Addendum to
S5.06, Bristol: Environment Agency.
European Commission Joint Research Centre, 2010. ILCD Handbook: Analysing of existing
Environmental Impact Assessment methodologies for use in Life Cycle Assessment.
European Union, 1999. Council Directive 1999/31/EC of 26 April 1999 on the landfill of waste.
European Union, 2008. Directive 2008/98/EC on waste.
European Union, 2010. Directive 2010/75/EU on Industrial Emissions.
Filippov, L. et al., 2009. Stabilisation of NaCl-containing cuttings wastes in cement concrete
by in situ formed mineral phases. Journal of Hazardous Materials, 171(1-3), pp. 731–8.
Fink, J., 2015. Water-based chemicals and technology for drilling, completion and workover
fluids. Waltham: Gulf Professional Publishing.
Flemming, C., 2000. A Discussion of Chemical Fixation and Solidification (CFS), Solidification/
Stabilization, Microencapsulation, and Macroencapsulation: API/NOIA Synthetic Based
Fluids Research Group.
Freidheim, J. & Candler, J., 2008. The Base Fluid Dilemma: What Can We Use and Where Can
We Use It?. Houston, Texas, AADE Fluids Conference and Exhibition.
GEA Group, undated. Slop Oil Recovery. [Online]
Available at: http://us.westfalia-separator.com/fileadmin/GEA_WS_US/Documents/
Brochures/Oil_Gas/GEA_Westfalia_Separator_Slop_Oil_Recovery.pdf
[Accessed March 2015].
Getliff, J. et al., 2002. Drilling Fluid Design and the use of Vermiculture for the Remediation of
Drill Cuttings. Houston, American Association of Drilling Engineers 2002 Technology
Conference “Drilling & Completion Fluids and Waste Management”.
Drilling waste management technology review 96

Gilbert, Hagstrom & Getliff, 2010. Reducing the Carbon Footprint of Drilling and Completion
Operations. Rio de Janeiro, SPE International Conference on Health, Safety and
Environment in Oil and Gas Exploration and Production.
Global Advantech, 2012. Cavitation Scrubbing of Solids and its Applications. [Online]
Available at: http://www.globaladvantech.com/Oilandgas/Oilandgas.htm
[Accessed March 2015].
Gonzales, M., Crawley, W. & Patton, D., 2008. New reduce, reuse recycle drilling waste
treatment technologies and programs. Houston, American Association of Drilling
Engineers.
Gumarov, S. M. et al., 2014. Drill Cuttings Reinjection Well Design and Completion: Best
Practices and Lessons Learned. Fort Worth, IADC/SPE Drilling Conference and
Exhibition.
Gumarov, S. M. et al., 2012. Drilling Waste Management. Oilfield Technology, December.
Guo, Q., Geehan, T. & Ovalle, A. P., 2007. Increased Assurance of Drill Cuttings Reinjection:
Challenges, Recent Advances and Case Studies. Drilling and Completion, pp. 99 - 105.
Guo, Q. & Nagel, N. B., 2009. Do’s and Don’ts in Drilling Waste Injection with Case Examples.
New Orleans, American Association of Drilling Engineers.
Halliburton, 2007. Baroid Surface Solutions™. ENVIRO-FIX™ Solids Stabilization Process. [Online]
Available at: http://www.halliburton.com/en-US/ps/baroid/drilling/waste-
management-solutions/waste-treatment-and-disposal/enviro-fix-solids-stabilization-
process.page
[Accessed 23 April 2015].
Halliburton, 2007. RotaVac™ Rotary Vacuum Dryer (RVD) Fluid Recovery and Cuttings Drying
System. [Online]
Available at: http://www.halliburton.com/en-US/ps/baroid/fluid-services/waste-
management-solutions/waste-handling-and-transport/rotavac.page
[Accessed March 2015].
Halliburton, 2014. Mobile Vertical Cuttings Dryer & Centrifuge Unit. [Online]
Available at: http://www.halliburton.com/public/bar/contents/data_sheets/web/
sales_data_sheets/h09978.pdf
[Accessed June 2015].
Halliburton, 2014. Vertical Cuttings Dryer & Centrifuge. [Online]
Available at: http://www.halliburton.com/public/bar/contents/Data_Sheets/web/
Sales_Data_Sheets/H010844.pdf
[Accessed March 2015].
Herigstad, T. P., Whyte, I., Kleppa, E. & Aas, N., 2010. Reduced Slop and Contaminated Mud in
Semi Submersible Rig Operations. Rio de Janeiro, Brazi, SPE International Conference
on Health, Safety and Environment in Oil and Gas Exploration and Production.
Hu, G., Li, J. & Zeng, G., 2013. Recent Development in the Treatment of Oily Sludge from
Petroleum Industry: A Review. Journal of Hazardous Materials, Volume 261, pp. 470–490.
Drilling waste management technology review 97

Ifeadi, C. N., 2004. The Treatment of Drill Cuttings Using Dispersion by Chemical Dispersion.
Port Harcourt, Nigeria, DPR Health, Safety & Environment (HSE) International
Conference on Oil and Gas Industry in Nigeria.
International Association of Drilling Contractors, 2010. Safety Alert – Cuttings Auger
Amputates Leg of Rig Crewmember.
International Association of Oil and Gas Producers, 2003. Environmental aspects of the use
and disposal of non-aqueous drilling fluids associated with offshore oil & gas operations.
International Finance Corporation, 2007. Environmental, Health and Safety Guidelines: Waste
Management Facilities.
International Finance Corporation, 2007. Environmental, Health, and Safety Guidelines for
Onshore Oil and Gas Development: World Bank Group.
IPIECA/IOGP, 2009. Drilling Fluids and Health Risk Management.
Ivan, C. & Dixit, R., 2006. Drilling Slop Water Reclamation. Houston, American Association of
Drilling Engineers Fluids Conference.
Ji, L., Guo, Q. & Wang, C., 2010. Is it Possible to Inject Larger Sizes and Higher Solids
Concentration of Drill Cuttings. Beijing, Society of Petroleum Engineers.
Kelland, M. A., 2014. Production Chemicals for the Oil and Gas Industry. Second Edition ed.
Boca Raton: CRC Press.
Khodja, M. et al., 2007. A Diagnostic of the Treatment of Oil Well Drilling Waste in Algerian
Fields. Recents Progres en Genie des Procedes, Volume Numero 94.
Koch Membrane Systems, 2013. Membrane Filtration Technology: Meeting Today’s Water
Treatment Challenges. [Online]
Available at: http://www.kochmembrane.com/PDFs/Other-Documents/Membrane-
Filtration-Technology---Koch-Membrane-Sys.aspx
[Accessed March 2015].
Kroken, A. et al., 2013. Evaluating an Alternate Cutting Separation Technology: Is the MudCube
a Fit for Brazilian Offshore Drilling?. Rio de Janeiro, Offshore Technology Conference.
Leonard, S. A. & Stegemann, J. A., 2010a. Stabilization/solidification of petroleum drill
cuttings. Journal of Hazardous Materials, Volume 174, pp. 463–472.
Leonard, S. A. & Stegemann, J. A., 2010. Stabilization/solidification of petroleum drill
cuttings: Leaching studies. Journal of Hazardous Materials, Volume 174, pp. 484–491.
Lloyds Register Energy, 2014. Understanding Offshore Container Certification: When, Where,
and Why. [Online]
Available at: http://www.lrqa.es/Images/26913-understanding-off-shore-container-
certification-wh.pdf
[Accessed June 2015].
Marin, J. U. et al., 2009. First Deepwater Well Successfully Drilled in Colombia With a
High-Performance Water-Based Fluid. Latin American and Caribbean Petroleum
Engineering Conference 31 May – 3 June, Cartagena de Indias, ColombiaCartagena de
Indias, Society of Petroleum Engineers.
Drilling waste management technology review 98

McCosh, K. & Getliff, J., 2004. Effect of Drilling Fluid Components on Composting and the
Consequences for Mud Formulation. Houston, AADE Drilling Fluids Conference.
McMillen, S., Smart, R. & Bernier, R., 2002. Biotreating E&P Wastes: Lessons Learnt from 1992
– 2002. Albuquerque, 9th Annual International Petroleum Environmental Conference.
MI-SWACO, 2010. Magna Centrifuge. [Online]
Available at: http://www.slb.com/services/miswaco/services/solids_control/
centrifuges/magna.aspx
[Accessed June 2015].
MI-SWACO, 2011. Offshore TCC Hammermill System. [Online]
Available at: http://www.slb.com/~/media/Files/miswaco/product_sheets/off_tcc_
hammermill.pdf
[Accessed March 2015].
MI-SWACO, 2013. The EnviroCenter Process. [Online]
Available at: www.slb.com/resources/other_resources/brochures/miswaco/
envirocenter.aspx
[Accessed March 2015].
MI-SWACO, 2014a. Vacuum Collection System (VCS). [Online]
Available at: http://www.slb.com/services/miswaco/services/cuttings_management/
cuttings_collections/vacuum_collection_system.aspx
[Accessed June 2015].
MI-SWACO, 2014. Screen Pulse Fluid and Cuttings Separator. [Online]
Available at: http://www.slb.com/~/media/Files/miswaco/brochures/screen_pulse_
brochure.pdf
[Accessed September 2015].
Morris , R. G. & Seaton, S., 2005. Unique Drilling-Waste Handling and Transport System
Offers Advantages to Drilling Operations and Assists Operators in Achieving Safety and
Environmental Compliance. Galveston, SPE/EPA/DOE Exploration and Production
Environmental Conference.
Morris, R. G. & Seaton, S., 2006. Design and Testing of Bulk Storage Tanks for Drill Cuttings
Offers Operators Safer Solution in Zero Discharge Operations. Houston, AADE Drilling
Fluids Technical Conference.
Mueller, F., Andrade, D. & Massam, K., 2013. Optimizing Drilling Waste Treatment to
Meet Discharge Criteria. Galveston, SPE Americas E&P Health, Safety, Security and
Environmental Conference.
Nagel, N. B., 2005. 4,000,000 Barrels and Counting: Experience with Cuttings Reinjection in
North Sea Shales. Anchorage, America Rock Mechanics Association.
Nahmad, D. et al., 2014. Treatment of Contaminated Synthetic Based Muds (SBM) Drilling
Waste in Block 47 Oman. Abu Dhabi, Society of Petroleum Engineers.
National Driller, 2015. Solid Answers to Questions About Drilling Fluids Disposal. [Online]
Available at: http://www.nationaldriller.com/articles/89886-solid-answers-to-
questions-about-drilling-fluids-disposal
[Accessed June 2015].
Drilling waste management technology review 99

National Petroleum Council, 2011. Paper #2–24 - Waste Management.


Nol-tec Systems, 2015. Dense Phase Pneumatic Conveying – A Gentle Way to Convey. [Online]
Available at: http://www.nol-tec.com/dense-phase-conveying.html
[Accessed September 2015].
NOV Brandt, 6th Edition. Handbook on Solids Control and Drilling Waste Management.
NOV, 2014. Dewatering Water-Based Drilling Fluids. [Online]
Available at: https://www.nov.com/Segments/Wellbore_Technologies/
WellSite_Services/Waste_Management/Drilling_Waste_Treatment/Dewatering_Water-
Based_Drilling_Fluids.aspx
[Accessed June 2015].
Ogechi Opete, S. E., Mangibo, I. A. & Iyagba, E., 2010. Stabilization/solidification of synthetic
Nigerian drill cuttings. African Journal of Environmental Science and Technology, 4(3),
pp. 149-153.
Ohio Environmental Protection Agency, 2012. Advisory Solidification and Disposal Activities
Associated with Drilling-Related Wastes at Solid Waste Landfills. [Online]
Available at: http://epa.ohio.gov/portals/34/document/NewsPDFs/Drilling%20
waste%20advisory%20final.091812.pdf
[Accessed 24 April 2015].
OSHA, undated. Oil and Gas Well Drilling and Servicing eTool. [Online]
Available at: https://www.osha.gov/SLTC/etools/oilandgas/drilling/drillingfluid.html
[Accessed June 2015].
Page, P. W. et al., 2003. SPE 80583: Options for the Recycling of Drill Cuttings. San Antonio,
Texas, SPE/EPA/DOE Exploration and Production Environmental Conference.
Paton, W. & Fletcher, P., 2008. Challenges of Waste Management in Environmentally
Developing Countries. Nice, SPE International Conference on Health, Safety and
Environment in Oil and Gas Exploration and Production.
Paulsen, Getliff & Sorheim, 2004. Vermicomposting and Best Available Technique for Oily
Drilling Waste Management in Environmentally Sensitive Areas. Calgary, Seventh SPE
International Conference on Health Safety and Environment in Oil and Gas Exploration
and Production.
Pereira, M. S. et al., 2013. Tests show potential of alternative method for offshore cuttings
drying. [Online]
Available at: http://www.drillingcontractor.org/tests-show-potential-of-alternative-
method-for-offshore-cuttings-drying-25578
[Accessed March 2015].
Petrobras, 2012a. 2011 Sustainability Report.
Railroad Commission of Texas, 2015. Summary Of Statewide Rule 8. [Online]
Available at: http://www.rrc.state.tx.us/oil-gas/applications-and-permits/
environmental-permit-types-information/swr8-summary/
[Accessed June 2015].
Drilling waste management technology review 100

Reddoch, J., 2000. Cuttings re-injection can solve disposal problems. Drilling Contractor,
July/August, pp. 28-29.
Rehm, B. et al., 2010. Underbalanced Drilling: Limits and Extremes. Houston: Gulf Publishing
Company.
Robinson, J. et al., 2010. Microwave Treatment of Oil Contaminated Drill Cuttings - Towards
a Commercial Scale System. Rio de Janeiro, SPE International Conference on Health,
Safety and Environment in Oil and Gas Exploration and Production.
Scomi Oiltools, 2008. Scomi Group Waste Management. [Online]
Available at: http://www.scomigroup.com.my/core/dwm/pdf/treatmentdisposal/
drill_cuttings_solidification_1208.pdf
[Accessed 27 April 2015].
Scomi Oiltools, undated. Vacuum Continuous Collection. [Online]
Available at: http://www.scomigroup.com.my/core/DWM/pdf/ContainmentHandling/
Vacuum_Cont_Collection_1208.pdf
[Accessed June 2015].
Seaton, S., Morris, R., Blonquist, J. & Hogan, B., 2006. Analysis of Drilling Fluid base Oil
Recovered from Drilling Waste by Thermal Desorption. San Antonio, 13th International
Petroleum Environmental Conference.
Soil Recovery A/S, n.d.. Incineration Plants. [Online]
[Accessed September 2015].
Stephenson, R. L. et al., 2004. Thermal Desorption of Oil from Oil-Based Drilling Fluids
Cuttings: Processes and Technologies. Perth, Society of Petroleum Engineers.
Storting White Paper No. 25, 2002-2003. The government’s environmental policy and the
environmental state of the realm. Norway: Storting.
Storting White Paper No. 58, 1996-1997. Environmental Policy for Sustainable Development.
Norway: Storting.
Storting White Paper No.38, 2003-2004. The Petroleum Activities. Norway: Storting.
Street, C. & Guigard, S., 2009. Treatment of Oil-Based Drilling Waste Using Supercritical
Carbon Dioxide. Journal of Canadian Petroleum Technology, Issue 06, pp. 26 - 29.
Svenson, T. & Taugbol, K., 2011. Drilling Waste Handling in Challenging Offshore Operations.
Moscow, Society of Petroleum Engineers.
TASC, 2015. Solidification | Liquid Waste Solution | Drilling Waste Solidification. [Online]
Available at: http://www.teamtasc.com/index.php/solidification-process
[Accessed 11 May 2015].
Tervita, undated. Caverns and Oilfield Waste Disposal. [Online]
Available at: http://www.tervita.com/solutions/challenge/waste-management-and-
disposal/caverns-and-oil-field-waste-disposal
[Accessed June 2015].
Drilling waste management technology review 101

The American Concrete Institute, 1999. Controlled Low Strength Materials (CLSM). Reported
by Committee 229. Report WSRC-TR-97-0100, Detroit, USA: The American Concrete
Institute.
Therm-tech, n.d. The TCC Process. [Online]
Available at: http://www.thermtech.no/Our-Technology/The-TCC-R-Process
[Accessed March 2015].
Trinity Environmental Services, 2015. Process. [Online]
Available at: http://trinityenv.com/process/
[Accessed June 2015].
US Environmental Protection Agency, 2015. Types of Composting. [Online]
Available at: http://www.epa.gov/composting/types.htm
[Accessed May 2015].
USEPA, 2000. Development Document for Final Effluent Limitations Guidelines and Standards
for Synthetic-Based Drilling Fluids and other Non-Aqueous Drilling Fluids in the Oil and
Gas Extraction Point Source Category.
USEPA, 2011. Information Sheet – Regulating Petroleum Industry Wastewater Discharges in
the United States and Norway.
USEPA, 2014. Proper Management of Oil and Gas Exploration and Production Waste. [Online]
Available at: http://www.epa.gov/epawaste/nonhaz/industrial/special/oil/hydrofrac.htm
[Accessed March 2015].
Visser, S., Lee, B., Fleece, T. & Sparkes, D., 2004. Degradation and Ecotoxicity of C14
Linear Alpha Olefin Drill Cuttings in the Laboratory and the Field. Calgary, Seventh
SPE International Conference on Health, Safety and the Environment in Oil and Gas
Exploration and Production.
Weatherford, 2012. OleoFoam Drilling Fluid System. [Online]
Available at: http://alshaheenweatherford.com/weatherford/groups/web/documents/
weatherfordcorp/wft132420.pdf
[Accessed October 2015].
Weatherford, 2014. Vertical Cuttings Dryer. [Online]
Available at: http://www.weatherford.com/doc/wft256026
[Accessed June 2015].
Western Oilfield Equipment Ltd, 2013. About the VAC-Screen Drilling System. [Online]
Available at: http://westoil.ca/vac-screen-4/about-vac-screen/
[Accessed May 2015].
Willis, J. M., Hester, M. W. & Shaffer, G. P., 2005. A mesocosm evaluation of processed drill
cuttings for wetland restoration. Ecological Engineering, 25(1), pp. 41–50.
World Oil, 2014. Drilling, Completion and Workover Fluids 2014, Houston: Gulf Publishing.
Yan, P. et al., 2011. Remediation of oil-based drill cuttings through a biosurfactant-based
washing followed by a biodegradation treatment. Bioresource Technology, 102(22),
pp. 10252–10259.
www.iogp.org
Registered Office Brussels Office Houston Office
City Tower Bd du Souverain,165 10777 Westheimer Road
40 Basinghall Street 4th Floor Suite 1100
14th Floor B-1160 Brussels Houston, Texas 77042
London EC2V 5DE Belgium United States
United Kingdom
T +44 (0)20 3763 9700 T +32 (0)2 566 9150 T +1 (713) 470 0315
F +44 (0)20 3763 9701 F +32 (0)2 566 9159 reception@iogp.org
reception@iogp.org reception@iogp.org

This report provides a review of the


drilling-specific wastes that are
generated during well construction
activities, both onshore and offshore.
This includes the following main types
of waste:
• drill cuttings and associated fluids, and
• interfacial mixtures.
The review focuses on waste types that
are specific to drilling operations, or
which present significant challenges
due to the quantities produced or
potential environmental impacts of
these wastes.
This report focuses on technologies,
methodologies, and processes for
managing waste once it has been
generated.

You might also like