technical briefing document for the task force on shale ... technical briefing document for the task...

157
1 Technical Briefing Document for the Task Force on Shale Gas Second Interim Report – Assessing the Impact of Shale Gas on the Local Environment and Health

Upload: buihuong

Post on 30-May-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

1

Technical Briefing Document for the

Task Force on Shale Gas Second

Interim Report – Assessing the

Impact of Shale Gas on the Local

Environment and Health

2

Contents

Introduction ............................................................................................................................................ 5

Hydraulic fracturing and the associated risks ......................................................................................... 6

Earthquakes ............................................................................................................................................ 9

Seismic activity in the UK .................................................................................................................... 9

Seismic activity caused by industrial activities. .................................................................................. 9

Seismic activity and hydraulic fracturing .......................................................................................... 13

Seismic activity monitoring techniques ............................................................................................ 16

The potential for subsidence due to shale gas extraction ................................................................ 19

Contamination ...................................................................................................................................... 21

Water ................................................................................................................................................ 21

Potential contaminants in water .................................................................................................. 22

Water usage during shale gas operations ..................................................................................... 24

Current U.S. National Energy Technology Laboratory research ............................................... 26

Ways that water contamination can take place ........................................................................... 27

Well integrity ................................................................................................................................. 28

Well installation methods ......................................................................................................... 31

Casing centralisation ............................................................................................................. 31

Variations in the cement used to construct wells ................................................................. 33

Corrosion of casing and degradation of cement ................................................................... 37

Well integrity testing ............................................................................................................. 38

Evidence of well integrity failure in the UK ............................................................................... 42

Evidence of well integrity failure in the US ............................................................................... 42

Current US National Energy Technology Laboratory research ................................................. 45

Evidence of water contamination ................................................................................................. 46

Water contamination associated with industrial activity ......................................................... 46

Water contamination associated with shale gas operations .................................................... 47

Contamination by upwards migration of fluids .................................................................... 50

Contamination by methane .................................................................................................. 56

Contamination of surface water ........................................................................................... 63

Groundwater contamination in the UK ........................................................................................ 66

Well abandonment ....................................................................................................................... 71

Waste residues .............................................................................................................................. 76

Air ...................................................................................................................................................... 82

3

Monitoring air quality and emissions ........................................................................................... 83

Leak and emission detection techniques .................................................................................. 83

Discrete ambient air measurements ........................................................................................ 85

Open source and whole site fence line monitoring .................................................................. 86

Mitigating and controlling emissions ............................................................................................ 88

Evidence of emissions ................................................................................................................... 92

Food issues associated with shale gas .............................................................................................. 96

Health issues associated with shale gas ............................................................................................... 97

Recent studies ................................................................................................................................... 99

The environmental health impacts review carried out by Werner et al. (2015) .............................. 99

Impact on water ............................................................................................................................ 99

Impact on health related air quality ........................................................................................... 100

Pollutants in soil .......................................................................................................................... 101

Occupational health .................................................................................................................... 102

Health impacts from infrastructure associated with shale gas operation .................................. 103

Social impacts.............................................................................................................................. 103

Public Health England 2013 Report ................................................................................................ 105

Air quality .................................................................................................................................... 106

Radon .......................................................................................................................................... 108

NORM .......................................................................................................................................... 110

Water and wastewater ............................................................................................................... 111

Hydraulic fracturing fluid ............................................................................................................ 113

New York State Department of Health 2014 report ....................................................................... 114

Air impacts .................................................................................................................................. 116

Water quality .............................................................................................................................. 116

Socioeconomic impacts............................................................................................................... 117

Health outcomes near sites ........................................................................................................ 117

High volume hydraulic fracturing health outcome studies ........................................................ 118

Birth outcomes ........................................................................................................................ 118

Case series and symptom reports ........................................................................................... 118

Local community impacts ....................................................................................................... 119

Cancer incidence ..................................................................................................................... 120

Shale gas Environmental Studies ................................................................................................ 120

Air quality impacts .................................................................................................................. 120

4

Water Quality impacts ............................................................................................................ 122

Induced earthquakes .............................................................................................................. 122

Conclusions from literature .................................................................................................... 123

Health Impact assessment .......................................................................................................... 123

Meetings with other States and consultation from medical professionals ................................ 124

Medact 2015 Report ....................................................................................................................... 125

Preese Hall, Lancashire, Case Study .................................................................................................... 131

The drilling process, well integrity and hydraulic fracturing .......................................................... 131

Fluid usage and waste disposal ....................................................................................................... 135

Seismic activity at Preese Hall ......................................................................................................... 136

Earthquakes resulting from hydraulic fracturing ............................................................................ 136

References .......................................................................................................................................... 138

Appendix 1 .......................................................................................................................................... 153

Appendix 2 .......................................................................................................................................... 156

5

Introduction

This briefing document was prepared for the Task Force on Shale Gas in order to inform and

ultimately to support the second interim report on environmental protection during shale gas

operations. The main aim of this document is to give an overview of the environmental

hazards that may arise from shale gas exploration and extraction in the UK. A great deal of

summary is presented concerning the presently in-place regulations for the minimisation of

hazards to the environment, and also of the new technical developments that are being made

to eliminate many of the problems that have become associated with the early years of shale

gas exploitation in the USA.

The document begins with a general introduction to hydraulic fracturing after which the main

environmental issues are discussed in turn. These are earthquakes, contamination (both

water and air) and health. The sections are arranged in such a way as to mirror the sections

in the second interim report. As part of the health section, a number of recent health studies

are discussed in detail. In addition, a short case study section on the Preese Hall site in

Lancashire is included at the end of the document.

6

Hydraulic fracturing and the associated hazards

The process of hydraulic fracturing (fracking) involves the injection of high pressure fluids

composed of water, proppant (sand particles used to keep open the fractures formed during

hydraulic fracturing) and chemicals into a geological formation (Figure 1). The aim of

hydraulic fracturing is to stimulate the flow of fluids (gas and formation water) from rocks of

low permeability (Healey, 2012). The fracking process over-pressurises the geological

formation which, in turn, results in the creation of clusters of new fractures in the rock and

the opening of pre-existing fractures. The fluid then flows into the fractures with the

proppant acting to hold the fractures open. Once the fracturing has been completed, the

pressure in the well is reduced which causes gas-bearing fluid to flow to the surface (Healey,

2012), pushing the frack-water up the borehole as flowback. The flowback is composed of

water with varying quantities of; proppant and chemicals used in the initial fracturing fluid;

gas and volatile organic compounds originating from the shale; minerals and metals in

solution (formation water); and naturally occurring radioactive materials (NORM)

(Department of Energy and Climate Change, 2014e). The NORM, minerals and metals

dissolved in the flowback are dissolved from the shale formations (Stamford and Azapagic,

2014). Once the gas has been extracted, it is processed on-site after which it can either be

sent for national distribution or it can be liquefied to produce liquefied natural gas (LNG)

which can be exported (Stamford and Azapagic, 2014). The life cycle of gas can be seen in

Figure 2.

7

Figure 1. Schematic diagram of the hydraulic fracturing process as carried out in the US. Not to scale. (From ProPublica, Granberg, 2015).

8

Figure 2. Simplified flow chart of the lifecycle of gas. The black boxes represent the stages that are unique to shale gas and the grey boxes represent stages unique to liquefied natural gas. (From Stamford and Azapagic, 2014).

The overall hydraulic fracturing operation, which includes the process of hydraulic fracturing

and associated industrial processes (e.g., water transport, well installation), pose potential

hazards to the environment. In geological terms, there is a likelihood of seismic activity as

large amounts of fluid are being injected into the Earth, thus modifying the in-situ stress

conditions resulting in the potential re-activation of faults (Healey, 2012). There is also

concern about the potential for the fracturing process to cause groundwater contamination

if the fractures intersect with permeable pathways (e.g., mineral veins or pre-existing fracture

networks) potentially allowing upwards migration of fluids (gas and possibly liquids) into less

deeply buried geological formation that might contain groundwater (Myers, 2012). Industrial

processes that can potentially impact upon the environment include the quantity and

composition of the fluids used during drilling and fracturing. Waste products (fluids, fugitive

emissions and drilling waste) also pose an environmental hazard if not properly treated and

disposed of. In addition, the integrity of the well is also a key concern as a loss of well integrity

can potentially lead to groundwater contamination (Jackson, 2014) and fugitive gas

emissions. Over the long term (years) the process of well abandonment becomes significant

because, even after it has been abandoned, the well can still pose a risk to the environment

through corrosion of any remaining casing or degradation of the cement, leading to well

integrity being compromised (Vengosh et al., 2014). There is the potential for groundwater

contamination and fugitive gas emission if well abandonment is not completed properly

(Osborn et al., 2011). The risk of adverse health effects associated with emissions, waste

products and groundwater contamination is also, sometimes, held to be of concern (Shonkoff

et al., 2014; Werner et al., 2015).

9

Earthquakes

Seismic activity can either be natural, i.e., movement of faults as a response to natural

changes in the in-situ stress state or of the mechanical properties of rocks in the Earth’s crust,

or can be induced, i.e., by human activities such as mining or fluid injection (Styles and Baptie,

2012).

Seismic activity in the UK

Seismic activity within the UK is relatively low when compared to other countries (The Royal

Society and the Royal Academy of Engineering, 2012). Earthquakes of magnitude ML4.0 (ML

means local magnitude) on the Richter Scale take place once approximately every 3-4 years

and events of magnitude 5.0 take place once every 20 years (The Royal Society and the Royal

Academy of Engineering, 2012). According to the British Geological Survey (BGS), who

operate and maintain a network of approximately 100 seismic monitoring stations (National

Earthquake Monitoring System) throughout the UK, the majority of the seismic activity is

around magnitude 1.5, which is at the detection limit of the BGS’ national monitoring system.

Hundreds of these earthquakes take place in the UK every year; however, on account of the

depth at which the earthquake takes place, very few are actually felt by the general

population. In essence, because of geometrical spreading (inverse-square law), the greater

the depth at which the earthquake takes place, the less chance there is of the effects being

felt at the ground surface, especially for small events.

Seismic activity caused by industrial activities

Davies et al. (2013) carried out an extensive review of seismic events induced by industrial

activities, these are summarised in Table 1 and Figure 3. The authors compiled evidence

dating from 1933 to the date of publication. They found 198 possible examples of induced

seismicity. However, they point out that, because they restricted their search to published

examples, their database cannot be considered comprehensive. They also do not account for

every seismic event in an earthquake swarm; they only report the largest magnitude event

during the swarm.

10

Table 1. Table displaying the types of industrial activities associated with earthquakes. The range of magnitudes associated with each activity is also shown. (From Davies et al., 2013).

Industrial activity Earthquake magnitude range

Mining 1.6-5.6

Oil and gas extraction 1.0-7.3

Water injection into oil wells 1.9-5.1

Waste disposal 2.5-5.3

Reservoir impoundment 2.0-7.9

Boreholes drilled by academic institutes 2.8-3.1

Solution mining 1.5-5.2

Geothermal operations 1.0-4.6

Hydraulic fracturing 1.0-3.8

Figure 3. Graph displaying the magnitude vs. frequency of felt seismic events induced by industrial activities. Note that the maximum magnitude of events associated with hydraulic fracturing is 4. (From Davies et al., 2013).

From examination of Table 1 and Figure 3, it is clear that the maximum magnitude of seismic

events associated with hydraulic fracturing are lower than those associated with the majority

11

of other industrial drilling / mining activities. It should be noted that there is no documented

evidence of seismic events originating from hydraulic fracturing being large enough to cause

structural damage or of inducing subsidence (Green et al., 2012). Of more significance is the

larger maximum seismic magnitude of events associated with waste fluid disposal into deep

geological formations. The Department of Energy and Climate Change’s (DECC) document on

water management during shale gas operations (Department of Energy and Climate Change,

2014e) presently makes no mention of whether this type of waste water disposal is permitted

in the UK.

The BGS has identified a multitude of instances of seismic activity being caused by other

industrial activities. Some of the best-documented examples are those associated with coal

mining in Nottinghamshire. Between mid-December 2013 and April 2014, 93 earthquakes of

maximum magnitude of 1.8 ML were detected around the New Ollerton area of

Nottinghamshire (Figure 4) (British Geological Survey, 2014). The area has a history of seismic

activity associated with local coal mining operations and the recent seismic events are also

considered to be related to this (British Geological Survey, 2014).

Figure 4. Map of the New Ollerton earthquake activity dating from 2000. (From British Geological Survey, 2014).

12

Historically, seismicity associated with abandoned coal mines in the UK has not exceeded a

magnitude of 3.0 ML (British Geological Survey, 2014). The maximum magnitude of seismic

events associated with shale gas activities are considered to be similar to that of coal mining

(Figure 3). However, the hypocentres (depths of origin) of coal mine related seismicity will

likely be shallower than shale gas related events. Therefore, a coal mining related event of

the same magnitude as a shale gas related event is more likely to be felt because even deep

coal mines only extend to half the depth that hydraulic fracturing of shale will extend.

Therefore, the same event will feel less strong from the greater depth on account of

geometric spreading. There are also no reports of mining related seismic activity resulting in

structural damage but there is evidence of superficial damage, e.g., cracks in plaster,

occurring as a result of mining-induced seismic activity (British Geological Survey, 2014).

Evidence from the US also indicates that seismic activity associated with coal mining is of a

magnitude similar to that of the maximum magnitude associated with shale gas operations.

For instance, Emery County, Utah, US, has a well-documented history of coal mining related

seismic activity (e.g., Arabasz et al., 2005; Fletcher and McGarr, 2005; McGarr and Fletcher,

2005). The maximum magnitude of such events is considered to be 3.9 ML (Arabasz et al.,

2005). This is comparable to the maximum magnitude of events associated with hydraulic

fracturing (3.8 ML) (Davies et al., 2013).

Seismicity is known to be induced by industrial activities other than coal mining. For instance,

the process of extracting geothermal energy has been documented as a cause for earthquakes

that are, potentially, larger than those caused by coal mining if the process takes place in

crystalline rock (igneous and metamorphic) as opposed to sedimentary rock (Evans et al.,

2012). One type of geothermal energy extraction, hot-dry-rock (HDR), involves the drilling of

two deep wells into hot regions of the Earth’s crust. The wells are then fractured and cold

water is pumped down through one of the wells. The water migrates through the fractures

connecting the two wells, during which time it takes heat from the fracture walls, and returns

to the surface at the other well. The process of fracturing and fluid migration can result in

seismic activity.

HDR geothermal energy exploration in the UK has been carried out at Rosemanowes Quarry,

near Penryn, Cornwall. The site, which was in operation between 1978 and 1991, originally

consisted of two wells drilled into granite to a depth of approximately 2 km and separated by

0.17 km, an additional well was drilled in 1985 (Evans et al., 2012). Local natural seismicity is

low, but there has been seismic activity associated with operation of the site. The largest

tremor was a 3.5 ML event which took place in 1981. This event was part of a cluster of

earthquakes that occurred 6 km south of the site, near the town of Constantine (Evans et al.,

2012). Initial fracturing using gel and water injection methods resulted in thousands of small

tremors of <0.16 ML, none of which were felt by the local population or site workers (Evans

et al., 2012).

13

In terms of the rest of Europe, the maximum recorded magnitude of earthquakes associated

with geothermal energy exploration and extraction from crystalline rocks is 3.5 ML (Evans et

al., 2012). This occurred at the Monte Amiata geothermal area of Italy where wells are drilled

to 3 km depth into metamorphic rocks (Evans et al., 2012). In Europe there has been some

effort to stimulate geothermal production from sedimentary rocks, however, very little in the

way of induced seismic activity has been documented. Of the activity that has been

documented, the largest recorded magnitudes have been 3.0 ML which occurred at

Larderello-Travale and Torre Alfina, both in Italy. The wells in each of these locations were

drilled to a depth of 2 km into carbonate rocks (Evans et al., 2012).

Seismic activity and hydraulic fracturing

Induced seismicity is generally accepted to occur when seven predefined criteria, as defined

by Davis and Frohlich (1993), are satisfied. These seven criteria are (Davies et al., 2013):

1. The seismic events are the first to be recorded in the area.

2. There is a time correlation between injection of fluid into the rock and the seismic

activity.

3. The epicentres are within 5km of the operation site.

4. The earthquakes occur at / around the depth of injection.

5. If tremors do not occur around the depth of injection, are there any geological

structures that may allow flow of fluid to the event hypocentre?

6. Are the changes in fluid pressure at the base of the well sufficient to permit

earthquakes?

In order for an earthquake to be induced due to hydraulic fracturing, a fault must be present

that satisfies three key factors: a) it must be already critically stressed; b) it must be able to

accommodate a large amount of the fracturing fluid (i.e. it must be sufficiently hydraulically

conductive for fluid injection into the fault to occur) and; c) it must be composed of rock

strong and brittle enough to allow seismic failure (Green et al., 2012).

Two types of seismic events are known to be associated with the process of hydraulic

fracturing (The Royal Society and the Royal Academy of Engineering, 2012). Smaller,

microseismic events are associated with the formation of new fractures, while movement

along pre-existing, pre-stressed faults can result in larger seismic events. In terms of scale,

microseismic events are faint events that cannot be felt on the Earth’s surface. The maximum

magnitude of the microseismic tremors is 0 on the moment magnitude scale; for comparison,

a felt earthquake has a moment magnitude on the order of 3 (Halliburton, 2011). These larger

earthquakes occur when the fracturing fluid migrates through the rock and along a pre-

14

existing fault and decreases the pressure holding the fault together, this allows the fault to

move, resulting in an earthquake. The magnitude (energy release) of the event is dependent

upon the size of the fault, the elastic stiffness of the rock and the amount by which the fault

slips. In the case of hydraulic fracturing, the stimulation is the volume of fluid that flows

through the fault.

Davies et al. (2013) proposed four potential pathways that would allow fluid to penetrate pre-

existing fractures (Figure 5). In circumstances where a fault can become re-activated, the

borehole does not necessarily have to intersect the fault in order for re-activation to take

place. The fault may be located hundreds of meters away from the well (Davies et al., 2013).

Based on this, it would be necessary for the operators to survey an area around the drilling

site by seismic reflection profiling with the aim of identifying faults that could potentially be

re-activated. There is no guarantee that such faults will always be detected. It would also be

necessary for the operators to have an understanding of the hydraulic conductivity of the

target formations and those surrounding it. If a large fault is present some distance away

from the furthest lateral extension of the borehole then understanding the hydraulic

conductivity of the surrounding formations might allow the operators to determine whether

fracturing fluids could migrate into, and along the fault.

Figure 5. Illustration of the potential mechanisms by which hydraulic fracturing fluid can infiltrate along pre-

existing faults. 1) Direct injection of fluids at pressure Pf along faults. 2) Flow of fluid through fractures created

during the hydraulic fracturing stage. 3) Flow of fluid along pre-existing fractures in the shale. 4) Flow of fluid

through permeable layers in the shale. (From Davies et al., 2013). (Note that it is much easier for fluid to

penetrate hydraulic fractures than fault planes critically oriented for slip, because of the greater value of

pressure, tending to keep such fault planes closed.)

15

In order to understand fully and evaluate the risk of earthquakes, the in-situ rock stresses and

the fracture network within the fracturing area must be established (Healey, 2012). The

operator is obliged by law to conduct a geophysical survey of a site for evidence of any large

faults (i.e., those that are visible on a seismic reflection profile) that, if stimulated, could

trigger an earthquake. With the resulting data, the operator can then take reasonable steps

to avoid any interaction with the fault throughout both the drilling and fracturing stages

(Department of Energy and Climate Change, 2014d).

Once drilling has been completed, the operator is prohibited from carrying out exploratory

fracturing until a series of small test hydraulic fractures are carried out. If these are successful,

and there is no evidence of enhanced seismic activity, the operator can begin the main

exploratory hydraulic fracturing stage (Department of Energy and Climate Change, 2014d).

DECC requires the installation of real-time seismic monitoring systems on-site (Department

of Energy and Climate Change, 2014b). DECC states that any seismic activity that occurs

during the operation of the well must be reported. The operator is responsible for this as well

as the mitigation and monitoring of any activity (Department of Energy and Climate Change,

2014b).

At the Preese Hall site (Cheshire, UK. Operated by Cuadrilla Ltd.) two small earthquakes took

place in April of 2011 (for more information, see the Preese Hall section of this report) the

causes of which were the subject of a number of investigations (de Pater and Baisch, 2011;

Green et al., 2012; Styles and Baptie, 2012). The investigations recommended the

introduction of a traffic light system to monitor and mitigate any potential risks posed by

earthquakes caused by hydraulic fracturing. This traffic light system has been adopted by

DECC (Table 2) with the limit for acceptable induced seismic activity being set at magnitude

0.5. This threshold equates to the normal background seismic activity caused by the passing

of trucks, trains or farming vehicles. It is above the magnitude frequently associated with

hydraulic fracturing and as such might act as a precursor to larger events (Department of

Energy and Climate Change, 2014d). Due to the lack of shale gas operations that have taken

place, there is no information regarding the practical implementation of this system, and it

likely that setting the limit for cessation of operations at ML = 0.5 is very conservative and may

be unfeasibly low. However, DECC have stated that they will keep up to date with new

research on the levels of seismic activity associated with hydraulic fracturing and adjust the

magnitude thresholds accordingly (Department of Energy and Climate Change, 2014d).

Table 2. Table displaying the traffic light system in place for shale gas exploration in the UK. (Adapted from Department of Energy and Climate Change, 2014d).

Traffic light colour

Magnitude

Action taken

Green <0.0 Regular operation of well.

16

Amber 0.0-0.5 Injection proceeds with caution. Potentially reduce injection

volume. Increase intensity of monitoring.

Red >0.5 Immediately stop injection and fracturing, bleed off the well

and continue monitoring.

The Preese Hall reports also suggested two mitigation methods that might be applied if

induced seismicity is detected. The first method is to limit the amount of fluid that enters the

shale and the injection rate after the fracturing has taken place. This involves an immediate

reduction of well pressure upon completion of the fracturing operation, thus allowing as

much fluid as possible to flow back to the surface. The second involves reducing the volume

of fluid used during the fracturing process (Green et al., 2012).

Seismic activity monitoring techniques

Detection of the number, and magnitude, of earthquakes depends on two main factors: the

proximity of the monitoring station to the epicentre of the seismic event and; the quality of

the monitoring equipment (Davies et al., 2013). Both factors will determine the number of

events and the lowest magnitude limit that can be detected (Davies et al., 2013). The closer

the monitoring station is to the epicentre, and the higher the quality of the detection

equipment, the larger will be the number of smaller tremors that can be detected. Increasing

the number and sensitivity of monitoring stations in close proximity to the source of the

events will increase the detection resolution. In the case of shale gas operations, this will

involve installing monitoring stations around the areas where hydraulic fracturing takes place.

In order to establish background levels of seismic activity around the operation site,

monitoring systems should be installed prior to any drilling taking place. It should be noted

that in order to constrain the location of an event, the same event must be detected by at

least three monitoring stations or preferably more detection at only one or two stations will

not be sufficient.

A significant point made by Davies et al. (2013) is that, due to the fact that the detection of

earthquakes is so dependent upon the placement and quality of the seismic monitoring

stations, comparisons of detected earthquakes between sites is very limited unless

monitoring systems and their placement are identical.

There are many methods of monitoring seismic activity. These include:

Tiltmeters – These are highly sensitive instruments that measure tilt (or rotation) of

the Earth’s surface near faults (United States Geological Survey, 2014). Electronic

tiltmeters are similar to a spirit level; they contain a compartment filled with

conductive fluid and a bubble (United States Geological Survey, 2009). Rotation of the

Earth’s surface causes the bubble to move, which provides a measure of the degree

of tilt (United States Geological Survey, 2009)

17

Microseismic – This technique involves the measurement of small-scale earthquakes

(events of <0.0 ML) that are of too low a magnitude to be felt at the surface (EGS

Solutions, 2015). This is also a passive technique; this means that it requires no seismic

vibration trucks or explosives that are commonly used in active seismic exploration

(Microseismic, 2014). The detectors (geophones or accelerometers) are arranged in

arrays across the area of interest. They can be installed on the surface, at shallow

burial depths or down monitoring wells (Microseismic, 2014). The installations can

either be permanent or temporary depending upon the needs of the operator. During

the hydraulic fracturing period, the technique allows the operators to map, both

horizontally and vertically, where the fractures are occurring underground. Any

seismic events that take place during the remainder of the operation can be detected

in the same way.

As with most issues relating to shale gas operations, the risks associated with induced

seismicity will vary from site to site. As a result, the type and density (number) of monitoring

stations will be varied from site to site depending on factors such as the level of background

seismicity. Despite the lack of shale gas exploration activity in the UK, some seismic

monitoring of sites has taken place (see the Preese Hall Case Study section) one example of

which is the monitoring carried out at the drilling site at Balcombe, West Sussex, operated by

Cuadrilla Resources Ltd. This monitoring was carried out by members of the University of

Bristol (Horleston et al., 2013). Four broadband seismometers were installed 1 month before

drilling took place (Figure 6). Monitoring was carried out prior to, and during, the drilling

operation. The objective of the monitoring was to establish background levels of seismic

activity and noise in addition to examining the detectability thresholds of the equipment.

18

Figure 6. Map displaying the location of the four monitoring stations (yellow pins, BA01-BA04) in relation to the Balcombe drilling site (red pin). Note that the fourth monitoring station (BA04) was located in close proximity to the approximately NW-SE trending train line. Image is about 1.5 km across. (From Horleston et al., 2013. Reproduced in accordance with Google fair use policy).

Data was continuously collected at all four stations from 1st of July through to 25th of

September 2013. This covered the period one month prior to drilling (in order to establish

the seismic background) and the full duration of the operation. The data was first examined

for seismic event triggers using a computer algorithm. A total of 134 were detected, however,

manual re-examination revealed that none could be classified as seismic events and that all

could be attributed to increased background noise (Horleston et al., 2013).

Results show that the largest and most significant source of seismic noise was that of train

movements originating from the London to Brighton rail line which was located 150m from

the fourth monitoring station (BA-04) (Figure 6) (Horleston et al., 2013). The level of

perceived vibration intensity produced by such train movements equates to the intensity of

a magnitude 1.5 earthquake originating at 3 km depth over a more prolonged period (note:

intensity describes how a tremor is perceived people and objects at the Earth’s surface,

19

whereas magnitude measures energy released at the source). The remaining sources of

seismic noise were other surface factors including cars, people, animal movements and

weather. However, these produce different frequencies and signal shapes that allow them

to be differentiated from earthquakes (Horleston et al., 2013).

When drilling began, the general level of background noise increased. As would be expected,

the monitoring station nearest to the drill site displayed the largest increase in background

(Horleston et al., 2013). In terms of detection limits, Horleston et al. (2013) estimate that the

minimum magnitude that could automatically be detected by the array and computer

algorithm was -0.2. The authors note that this is slightly lower than the limit of what is

required for implementation of the traffic light monitoring scheme.

The report highlights the importance of appropriate monitoring site selection. Sites need to

be located at an appropriate range of distances from the drilling site whilst at the same time

avoiding pre-existing noise. For example, the similarity between felt intensities at the surface

arising from a 0.5 ML earthquake occurring at depth and train movements on the surface

could result in the masking of a seismic event if it were to take place as a train was passing

(Horleston et al., 2013). Although this is unlikely, it is still a point worth making as a tremor

of this magnitude would cause the halting of any operation under the current traffic light

monitoring system. Horleston et al. (2013) also recommend that a real-time monitoring

system be installed at each site. Because of the need for rapid analysis of seismic data and

the need to differentiate between background noise and seismic events, the authors raise the

point that, based on their lower detection limit being near to that required under the

proposed traffic light system, the UK may not currently be in a position accurately to predict

seismic events rapidly enough to allow hydrofracturing to be halted before more events take

place. The fact that the felt intensity arising from train movements, albeit detected at a

monitoring station near to a train track, can be larger than that required to halt fracturing

demonstrates this point.

According to the authors, the UK’s recording of seismic events by the BGS is only complete to

a minimum magnitude of 2.0 ML, or possibly 1.5. Using the Gutenberg-Richter relationship

for the UK, the authors estimate that there is the potential for 5,000 natural earthquakes per

year in the UK that would trigger an amber warning and 2,000 that would trigger a red

warning. This highlights the need for thorough baseline monitoring prior to any fracturing

taking place, the difficulty of recognising induced seismicity at this level relative to

background and the potential need for revision of the limits set in the traffic light system as

experience is acquired.

The potential for subsidence due to shale gas extraction

DECC (Department of Energy and Climate Change, 2014d) consider that there is minimal risk

of subsidence resulting from the natural gas extraction process. Even after the drilling of

20

hundreds of thousands of wells in the US, there is currently no documented evidence of

subsidence resulting from shale gas extraction (Department of Energy and Climate Change,

2014d). On the other hand, subsidence has been associated with coal mining activities due

to the fact that large amounts of material are removed from the subsurface (Singh and Yadav,

1995). Shale gas operations will remove an extremely small volume of material from the

subsurface, therefore the risk of subsidence can be considered negligible. Subsidence can

also happen when sufficiently porous rock is compressed at high enough pressures to begin

to collapse the porosity, but shale is a low-porosity rock that is not easily compressed and will

therefore be unable to collapse and cause subsidence.

21

Contamination

Water

A concern amongst the general public is the perceived potential for water contamination to

occur as a result of hydraulic fracturing. Ground and surface water contamination is

intimately linked with other aspects, including preparation of the site, i.e., the drilling phase,

well integrity problems, well abandonment and management of waste residues and the risk

they pose for water contamination.

Before proceeding with the discussion of water contamination and shale gas operations, it is

worth considering exactly what an aquifer is. An aquifer is simply a permeable water-bearing

rock formation, regardless of whether or not it is exploited for potable water. Most UK

aquifers are not exploited but potable water is drawn from rivers and surface reservoirs

instead. Two thirds of the UK domestic water supply is drawn from surface water sources.

It should also be noted that drilling through aquifers is not an uncommon occurrence in the

UK, as shown in Figure 7 (Davies et al., 2014). Between 1902 and 2013, 2152 hydrocarbon

wells have been drilled in the UK. 428 (20%) have been drilled through highly productive

aquifers and 535 (25%) were drilled through moderately productive aquifers (Davies et al.,

2014).

22

Figure 7. Map of the UK showing the location of intergranular flow and fracture flow aquifers together with a) the location of onshore exploration wells, and b) potential shale gas and oil-bearing formations. (From Davies et al., 2014).

Potential contaminants in water

A variety of chemicals are used in the hydraulic fracturing fluid. The most common additives

are shown in Table 3. The amounts and types of chemicals added to the fluid varies from site

to site. However, in all cases, the aim is to optimise the hydraulic fracturing process and

maximise gas recovery. In the US, the main additives are friction reducers (polyacrylamide)

designed to allow the fracturing fluid to be pumped into the well at an increased rate; this is

known as “slickwater” fracturing.

23

Table 3. Table displaying the chemicals generally used in the hydraulic fracturing fluids together with their purpose and downhole results. (From Frac Focus, 2015b).

The chemicals in the fluid are subject to assessment by the appropriate environmental

agency. The details of the chemicals, together with the reason for their use and associated

hazards, must be fully disclosed. This is subject to the protection of intellectual property of

the operators (Department of Energy and Climate Change, 2014e). However, it is in the

interests of the operators to publically disclose all chemicals as it helps gain public trust and

aids transparency. In the US, a chemical disclosure registry (www.fracfocus.org) has been set

up. Members of the public can look up any registered well and find the composition of the

fluid used during the fracturing process together with information on well depth and water

volume used. A similar, centralised, platform has been put in place in Europe by the

International Association of Oil and Gas Producers (http://www.ngsfacts.org/findawell/). The

website lists the shale gas wells that have been fractured since 2011 by operators that

participate in the Natural Gas from Shale Facts website (http://www.ngsfacts.org/), a website

providing factual information on hydraulic fracturing). The website provides the location,

permitting information and the substances used in the fracturing process. However,

disclosure of information about wells is voluntary.

There is some debate around the composition of the hydraulic fracturing fluids, mostly

concerning the chemical additives. For instance, a report by the Tyndall Centre for Climate

Change put the typical chemical content of the fluid at 2 vol%; this translates to 180-580 m3

24

of chemicals per well mixed into the fracturing fluid (Wood et al., 2011). This concentration

is higher than the estimates suggested in the Royal Society report (0.17 vol%) (Stamford and

Azapagic, 2014). The amount of chemicals introduced into the well, based on this estimate,

is 4.5-14.5 m3. It should be noted that current drilling in the UK is still at the exploratory stage

and the composition of the fluid may change if the operators move into the production stage

(Stamford and Azapagic, 2014). It should also be noted that the composition of the hydraulic

fracturing fluid will vary from site to site on account of factors such as variations in local

geology, well depth and well length.

Water usage during shale gas operations

The process of drilling and hydraulic fracturing consumes large amounts of clean water which

must be sourced by the operator either from a local utilities company or from local

ground / surface water sources. Before utility companies provide water to the drilling site,

they must be satisfied that supplying the requested amount of water will not put domestic

water (and other customers’) supplies at risk (Department of Energy and Climate Change,

2014e). Nonetheless, it should be noted that this is no different to other industries which

require water. The permission to extract water from nearby surface water or groundwater

sources is dependent upon the operator obtaining a permit to do so from the appropriate

environmental agency. The main criteria that must be met in order for a licence to be granted

is that the water supply in the area must be sustainable (Department of Energy and Climate

Change, 2014e), it must also not impact adversely on other users and the environment. A

further factor that needs to be considered is the exact time when the water will be needed

on-site. The operation will not require a constant supply of water, but rather will require

larger amounts of water at particular times, i.e., during drilling and at the start of the hydraulic

fracturing stage (The Royal Society and the Royal Academy of Engineering, 2012).

The amount of water required to conduct hydraulic fracturing is considerable, although it is

not as water intensive as some other industries, e.g., beverage and food production, and

paper production (Department of Energy and Climate Change, 2014e). It is estimated that

each well will require between 10,000 and 30,000 m3 (10,000 to 30,000 tonnes or 2 to 6

million gallons) of water over the course of the operation (Logan et al., 2012). Most of this

water will be required in the drilling and hydraulic fracturing stages, i.e., 1-2 months. UK

operator, Cuadrilla, estimates that 12,000 m3 of water will be required over the life span of a

well (House of Commons Energy and Climate Change Committee, 2011). This equates to the

same amount of water required to run a coal-fired power station for 12 hours, or to water a

golf course for one month (The Royal Society and the Royal Academy of Engineering, 2012),

or the amount lost each hour by United Utilities from leakages.

Vengosh et al. (2014) pointed out that a potential way to reduce the impact on the domestic

water supply would be to use alternative water sources as a base for the drilling mud and

hydraulic fracturing fluid. The authors suggested that low quality water, such as brackish

25

water or water from acid mine drainage, could be used. However it is uncertain whether such

water would be available in the quantities needed for hydraulic fracturing. There is evidence

that when acid mine drainage waters are mixed with Marcellus Shale flowback, various

dissolved salts are formed that can act to capture contaminants within both fluids (Kondash

et al., 2014). However, the effect that these already contaminated waters might have on the

physical properties of the fracturing fluid is unknown. Perhaps an unreasonable amount of

chemicals may need to be added to the fluid in order for it to have the same properties as

fracturing fluid prepared with uncontaminated water as the base. This could offset the

environmental benefits of reducing the stress on the domestic water supply.

During the drilling stage a fluid known as “drilling mud” is permanently circulated through the

borehole. The fluid is pumped down the drill string and exits the drill bit, at high pressure,

through nozzles. It then travels back to the surface around the gap between the drill string

and the wall rock, known as the annulus. This process acts to lubricate and cool the drill bit

whilst loosening and collecting fragments of rock resulting from the drilling, known as

“cuttings” (Williamson, 2013). The drilling mud transports the cuttings to the surface, thus

allowing the drill bit to function properly and not become clogged up. In addition, a powdered

mineral, barium sulphate (barite), is added to the drilling mud to increase the density of the

fluid so that the hydrostatic pressure created by the mud column is greater than the reservoir

pressure. This prevents hydrocarbons from the reservoir from flowing into the borehole

(Williamson, 2013). The hydrostatic pressure also helps to stabilise the borehole, by

counteracting the forces due to depth of burial that make the borehole want to close up.

Once returned to the surface, the fluid passes through a shaker that separates the larger

cuttings from the drilling mud. The mud then passes through a series of tanks that remove

the smaller cuttings through hydrocylones or centrifuges and apply chemical treatment to

maintain the desired specifications, thus allowing the fluid to be recycled (Williamson, 2013).

The most commonly used fracturing fluid consists of a water base (approximately 90 to 93%)

together with sand proppant (approximately 5 to 8% by volume) and chemicals (1-2%). The

base can also be foam, oil or acid. Which of these is chosen depends upon the depth to the

formation of interest. For instance, shallow, low-pressure reservoirs can be fractured with

foams created using N2 or CO2 while acids are used in reservoirs mainly consisting of

carbonate rocks (Holditch, 2006). Oil-based fluids are used when fracturing takes place in

close proximity to formations that are sensitive to water damage (Montgomery, 2013).

However, depending on the oil used, there is the potential for the flowback to be more highly

contaminated and hence require more treatment before being disposed of. However, it is

unlikely that an oil-based fracturing fluid would be permitted in the UK.

Vidic et al. (2013) highlighted concerns around the fate of hydraulic fracturing fluid. Two

major questions they raised based on their literature review were; 1) what happens to the

water that does not return to the surface as flowback or production water, and 2) could the

non-recovered water eventually contaminate aquifers (Davies, 2011; DiGiulio et al., 2011;

26

Boyer et al., 2012; Myers, 2012; Warner et al., 2012). One potential fate of the fracturing

fluid is that it is absorbed into the shale. A previous study of Marcellus Shale well logs has

shown that the rock contains little free water (Engelder, 2012), therefore there is the

potential for the fluid to enter the rock. There is also potential for fluid to migrate upwards

along gaps between the casing and cement or along fractures in the rock formations, although

this requires a suitable pressure gradient.

Current US National Energy Technology Laboratory research

The US Department of Energy (DOE) and the US National Energy Technology Laboratory

(NETL) are currently funding three projects examining alternative, non-water-based

stimulation techniques. The projects are as follows (National Energy Technology Laboratory,

2015e);

Development and field testing of novel natural gas surface process equipment for

replacement of water as primary hydraulic fracturing fluid. This project involves the

use of wellhead (produced) natural gas that has been liquefied and compressed, as

the primary constituent of hydraulic fracturing fluid. If successful, this may prove to

be a more cost-efficient non-water / CO2 based stimulation technology that can be

used instead of, or in parallel with, traditional water-based methods. The benefits of

such fracturing fluids include less waste production, reduced need for water transport

and better production through reduced clay swelling and blockages in the producing

formation. The project is due to finish in October 2017 (National Energy Technology

Laboratory, 2015a).

Development of nanoparticle-stabilized foams to improve performance of water-less

hydraulic fracturing. This project aims to develop surface-treated nanoparticles

capable of stabilizing foam fracturing fluids, the main constituents of which are

commonly used in fracturing fluids (CO2, N2, water and liquefied petroleum gas).

Nanoparticles treated with surface coatings could provide long term stability for

foamed fracturing fluids. This, in turn, can reduce water usage. In addition, their small

size means that they can stabilize foams with much smaller bubble sizes, and hence

permit an increase in the viscosity of the foam. This has the added benefit of allowing

proppant to be carried in the foam. Another benefit of using nanoparticles is that the

type of treatment and concentration of the particles themselves can be tuned to a

particular system. For example, the particles can be tuned to allow the proppant to

be carried into the fractures at which point the foam structure will break, this

facilitates flowback without re-foaming of the fluid. The project is due to finish in

October 2016 (National Energy Technology Laboratory, 2014).

Development of non-contaminating cryogenic fracturing technology for shale and

tight gas reservoirs. The project aims to study, test and develop cryogenic fracturing

technology using CO2 or liquid nitrogen. If successful, this research will increase the

27

permeability of, and therefore recovery from, shale gas reservoirs. In addition to

increasing production, this technique can reduce, or even eliminate, water usage,

whilst also reducing damage to the target formation and reducing any potential for

groundwater contamination by eliminating the need for additives in the fracturing

fluid. The project is due to end in July 2016 (Research Partnership to Secure Energy

for America, 2013).

Ways that water contamination can take place

The Groundwater Foundation, a US-based group, cite a number of potential mechanisms of

groundwater contamination. These include leaks from storage tanks, septic systems, landfills,

chemicals and road salts (The Groundwater Foundation, 2015). When considering the

potential for water contamination from shale gas activities, the contamination risk from these

other sources must be considered in order to provide a perspective on the severity of any

hazard posed and the associated risks to humans and animals.

In the shale gas industry, storage tanks are used on-site to store water and waste fluids.

Outside of the shale gas industry, storage tanks can be located both above and below ground

and can contain a variety of fluids, including other hydrocarbons. Over time, and without

proper inspection and repair, these can leak and potentially result in contamination of surface

and / or groundwater (The Groundwater Foundation, 2015). The Groundwater Foundation

say that there are as many as 10 million storage tanks in the US, these are likely to be for both

commercial (e.g., fuel storage at petrol stations and chemical storage at factories) and private

use (e.g., on farms). One would presume that those at commercial sites would be subject to

regular checks to ensure that no cracks are appearing, however, this may not be the case for

storage tanks in private use. This could potentially increase the risk of leaks, and hence of

water contamination. In terms of shale gas operations, on-site storage tanks will fall under

the first category and should be periodically inspected. This should reduce the hazard to

acceptable levels. To mitigate this hazard further, drilling and fracking sites are routinely

covered with geotextile membranes designed to contain any spills or leakages. This was

observed to be the case at the US sites visited by the Task Force in March of 2015. Bunds and

additional ground membrane protection can be used in areas of chemical storage to mitigate

further against spills.

Septic tanks, particularly those on private properties not connected to the main sewage

system, can also present a significant risk to surface and groundwater if not properly

designed, constructed, located or maintained (The Groundwater Foundation, 2015). A leak

from one of these tanks could result in the discharge of human waste, and the chemicals used

to treat it, into the local groundwater system.

Landfill sites have previously been shown to have the potential to cause groundwater

contamination through migration of leachates (meteoric fluid (i.e. rainwater) that percolates

28

through waste and leaches contaminants) (e.g., Barker et al., 1988; Monteiro Santos et al.,

2006; Mor et al., 2006). Modern landfills are sealed against leaks by first laying down a layer

of low permeability clay. A synthetic membrane layer is then placed over the top of the clay.

This is an impermeable layer that prevents any leachates migrating into the surrounding rock

and causing contamination. These linings can, with time, degrade and crack, thus allowing

leachates to move into the surrounding clay and increasing the risk of contamination.

Chemicals, such as fertilisers and pesticides used on farmland, and road salts also have the

potential to contaminate surface water and groundwater (e.g., Levallois et al., 1998; Babiker

et al., 2004). The process of eutrophication occurs when excessive quantities of chemical

fertilizers are used on crops. These fertilizers can enter surface waters via runoff or can

percolate through the soils and can enter into shallow aquifers. Once in surface waters, the

chemicals encourage the growth of algae on the water’s surface. This results in the oxygen

content of the water decreasing and the hazard to aquatic species increasing. Salts used to

de-ice roads during cold periods can migrate into the groundwater system, resulting in

elevated chloride levels (e.g., Williams et al., 2000; Godwin et al., 2003). This is a particular

problem in areas that require regular de-icing during cold periods, i.e., urban areas and

motorways (The Groundwater Foundation, 2015).

Well integrity Wells consist of a number of barriers, i.e., casing, cement, valves and seals, which prevent the

unplanned escape of fluid from the well (Davies et al., 2014). In the UK, DECC state that a

well must have at least three layers of casing; an outer conductor or surface casing, an

intermediate casing that extends down below the aquifer and an inner production casing

which runs into the geological formation of interest (Figure 8) (Department of Energy and

Climate Change, 2014c). It is this latter section in which the hydraulic fracturing takes place.

For information on the drilling and fracturing process at Preese Hall, see the Case Study

section. The operator can add further layers of casing to improve the well stability and further

reduce risks of well leakage.

Well integrity refers to the isolation of gas originating from the target formation and the

formations through which the well passes (Jackson, 2014). A failure in well integrity involves

the failure of one or more barriers that leads to the formation of a pathway that allows

leakage of liquids and / or gases from the well into the surrounding environment (King and

King, 2013) or along the outer wall of the well casing. It is worth noting that a barrier failure

will not necessarily lead to a failure in well integrity as a complete pathway may not be

formed. Only failure of all barriers will result in a leak path forming. A barrier failure is one

that does not lead to complete loss of integrity. For instance, a single layer of casing can fail

but none of the fluid within the well may leak out into the surrounding rocks, therefore,

overall well integrity is maintained.

29

Figure 8. Schematic diagram of the design of a shale gas well. Not to scale. (From Stamford and Azapagic, 2014).

Leaks require a fluid source, a driving force for fluid motion and failure of one or more of the

barriers (Davies et al., 2014). Potential fluid sources are shown in Figure 9 and potential flow

pathways are shown in Figure 10. The fluid source is most commonly the drilling or fracturing

fluid. Driving forces arise due to fluid pressure gradients or fluid buoyancy (Figure 9).

Pathways can form due to cracking of cement, cement shrinkage, high cement permeability

and the cement failing to fill pores in the wall rock and irregularities in the borehole wall. In

extreme cases, events such as rock movement can cause the casing to buckle and / or shear

(Figure 10) (Davies et al., 2014). Three of the more common modes of well failure are: (1)

blowout due to the uncontrolled escape of fluid from the well; (2) annular leaks where the

30

fluids can move upwards along the outside of the outer well casing or gaps within the layers

of the well casing and radial leaks; and (3) where the well casing fails and the fluid leaks into

the surrounding rocks (Figure 9) (The Royal Society and the Royal Academy of Engineering,

2012).

Figure 9. Schematic diagram of the potential sources of fluid that can enter the well casing if a failure of well integrity takes place. 1) Gas-rich coal formations. 2) Non-producing permeable gas reservoir. 3) Biogenic or thermogenic gas in a shallow aquifer. 4) Gas from the target reservoir. (From Davies et al., 2014).

31

Figure 10. Schematic diagram displaying the potential pathways along which fluid can exit the well, resulting in leaks. 1) Through the annulus between cement and wall rock. 2) Between layers of casing and cement. 3) Between the cement plug and casing. 4) Through the cement plug. 5) Through the cement. 6) Through the cement then along the boundary between the cement and the casing. 7) Along the plane of weakness caused by shearing of the wellbore. (From Davies et al., 2014).

Well installation methods

The following section outlines the steps taken by operators during installation to ensure that

well integrity is established and maintained.

Casing centralisation

Effective centralisation of the casing is critical to achieving good cementation. The correct

installation of both allows the casing to be properly supported; prevents fluids from leaking

to the surface; and allows production zones and water bearing zones to be properly isolated

(PetroWiki, 2014a). There are two common methods of centralising casing during the

installation process; the use of (a) bow-spring and (b) rigid type centralisers (Figure 11). The

former method is the most popular and can be used when the borehole is enlarged. On the

other hand, if the borehole is only slightly larger in diameter than the casing, rigid centralizers

32

are used. Rigid centralizers are also frequently used when installing the casing in the

horizontal sections of wells (PetroWiki, 2014c).

Figure 11. Schematic diagrams (not to scale) of the two main centraliser types used in the oil and gas industry. Left: Rigid centralizer. Right: Bow-spring centralizers. Note that the dimensions of the centralizers will vary depending on which casing layer is being installed. (From Halliburton, 2006a, b).

Effective casing centralisation is also crucial for ensuring well integrity as the

non-centralisation of casing within the borehole can leave an area of un-cemented casing if

mud is not properly displaced during the cementing process (Figure 12 top right). This sort of

gap should be routinely detected by the operators (see Well Integrity Testing section below)

and remedial action should take place. If, however, the re-cementing is not carried out

properly, fluid can be trapped in the annulus and interact directly with the casing. This will

likely result in severe corrosion and an increase in the risk of well failure (Choi et al., 2013).

33

Figure 12. Schematic cross-sectional diagrams of some potential mechanisms of corrosion and subsequent well integrity loss. Top left: Idealised well with no loss of well or barrier integrity. Top right: Off centred casing resulting in the formation of a large annulus on one side of the borehole. Bottom left: Annulus formed between layers of casing. Bottom right: Stress cracking takes place in the cement (From Choi et al., 2013).

Variations in the cement used to construct wells

The most commonly used cement in the oil and gas industry is referred to as APO Oil Well

Cement (Environment Agency, 2012b). The cement is almost exclusively Portland cement

which is a burned mix of limestone and clay (PetroWiki, 2014b). This particular type of cement

is used as it is readily available, its physical properties also make it applicable to a wide range

of drilling operations, e.g., it can easily be pumped into boreholes and it sets easily (even

underwater) (PetroWiki, 2014b).

Another factor that makes Portland cement applicable to many different drilling scenarios is

that it can be easily modified (PetroWiki, 2014b). There are two main types / classifications

34

of Portland cement that are produced; these are the American Society for Testing Materials

(ATSM) and American Petroleum Institute (API) (PetroWiki, 2014b). The ATSM cement

classification is mainly used in the construction industry, and therefore will not be considered

further. The API classification, on the other hand, applies only to cement used in the

construction of oil and gas wells. The API cement classification has a number of sub-classes

dependent upon the chemical composition of the slurry. API cement can be Class A through

to J (excluding I) with classes G and H being the most commonly used (Table 4) (Crook, 2006;

PetroWiki, 2014b). The composition of typical Class G and H Portland cements can be seen in

Table 4, a comparison between carbon compounds in different cement classes and the effect

that varying the amount of specific compounds makes, is shown in Table 5. More detailed

information about the chemical composition and material properties of the cements can be

found in Appendix 1. Based on the information presented in the below tables, it is clear that

the optimum cement will vary from site to site.

Table 4. Table displaying the depth and special conditions that dictate which API Class of cement can be used in a given well. (From King, 1996).

API Class

Depth to base of well (m) Special conditions for use

A 1830 None

B 1830 Sulphate resistance required

C 1830 Finer grind giving high early strength

D 1830-3050 High pressure and temperature conditions

E 3050-4270 High pressure and temperature conditions

F 3050-4880 Extremely high temperatures

G 2440 Can be mixed with additives and used over a range of temperatures and depths

H 2440 Can be mixed with additives and used over a range of temperatures and depths

J 3600-4880 High pressure and temperature conditions and can be mixed with additives

Table 5. Table displaying the carbon compound content of the various API Classes of cement. The effect of varying the carbon compound content is also shown. Note that under the phase composition section, C is an

35

abbreviation of CaO; S is and abbreviation of SiO2; A is an abbreviation of Al2O3; and F is an abbreviation of Fe2O2. (From Crook, 2006).

For the cement to be pumped to the base of the well, i.e., >3 km distance, additives are

introduced into the slurry in order to control the density which, in turn, controls the rock

formation pressure during setting, setting time, flow properties, and when set, the strength

(Environment Agency, 2012b). This also represents one of the reasons why Class G and H

cements are the most popular in the oil and gas industry as their physical properties can be

easily modified through the introduction of such additives (Crook, 2006).

Accelerators, e.g., CaCl2, NaCl, KCl and Na2SiO3 are one type of additive. These act to alter the

time required for the cement to harden and set. These are particularly useful in areas of low

temperature country rock where adding an accelerator will result in a more efficient drilling

and cementing phase (Crook, 2006).

Retarders are most commonly used in the Class A, C, G and H cements. These serve the

purpose of extending the thickening time of the cement. Note that the thickening time is the

time required to mix and pump the cement slurry. Examples of retarders include

36

lignosulfonates, cellulose derivates, hydroxycarboxylic acids, organophosphates and

inorganic compounds (e.g., borax (Na2B4O7∙102) and ZnO) (Crook, 2006).

Lightweight additives, also known as extenders, can be added to API Class A, C, G and H

cements. When the cement slurries are mixed to these API Classes, the resulting slurry can

be too dense (i.e., >15 lbm/gal (pound mass per US gallon)) to allow efficient circulation

(Crook, 2006). The addition of extenders solves this problem by reducing the density of the

cement (Crook, 2006).

There are three common types of extenders; physical, pozzolanic and chemical (Crook, 2006).

One common physical extender is bentonite gel. Bentonite is a colloidal clay (i.e. the mineral

group including montmorillonite [NaAl2(AlSi3O10)∙2OH]). Attapulgite

[Mg,Al)2(OH/Si4O10)∙12H2O], a salt gel, is also used in slurries with high salt content.

Attapulgite is a mineral with a fibrous crystal habit (shape) not dissimilar to asbestos. Because

of this, its use has been banned in some countries but not in the UK. Perlite (silica rich volcanic

glass), crushed coal and ground rubber are also used (Crook, 2006). One or more of these

extenders may be added to any one slurry.

Pozzolanic extenders have a lower specific gravity than the standard API Class slurries;

therefore, by adding them to the mix, the density of the slurry is reduced whilst the

consistency of the slurry remains approximately unaltered (Crook, 2006). Four main types of

pozzolanic extenders are used; fly ash (a mix of SiO2+Al2O3+Fe2O3), microspheres (hollow, gas

filled silica rich aluminosilicate glass spheres), microsilica (high surface-area vitreous silica and

SiO2 mix) and diatomaceous earth (diatom skeletons) (Crook, 2006).

Gypsum (hydrated calcium sulphate) and sodium silicate are the two most commonly used

chemical extenders. The latter is much more effective at lowering the slurry density than

other extenders, particularly bentonite, therefore less needs to be used (Crook, 2006).

Weighting agents do the opposite of extenders, i.e., increasing the density of the slurry

(Crook, 2006). This is required in situations where wells are highly pressurised and require a

slurry of higher density. Haematite (Fe2O3), ilmenite (FeOTiO2), hausmannite (Mn3O4) and

barite (BaSO4) are the main weighting agents used, with haematite being the most common

(Crook, 2006).

Dispersants, or friction reducers, are used to control the flow properties of the slurry and

reduce frictional pressure during pumping. The two main dispersants are polyunsulfonated

naphthalene and hydroxycarboxylic acids (i.e., citric acid) (Crook, 2006).

Fluid-loss-control additives (FLAs) are used to maintain the fluid volume of the slurry (Crook,

2006). This is significant as many of the other physical properties of the slurry depend upon

the water content. Using FLAs ensures that the water content, and hence physical properties,

remain consistent. The materials used as FLAs can be broadly categorised as either water-

soluble or water-insoluble. Water-insoluble FLAs are either bentonite or polymer resins while

37

water-soluble FLAs can be natural polymers, cellulosics or vinylinic-based polymers (Crook,

2006).

Also of note is the fact that cement slurries can be “foamed” through the introduction of a

foaming agent, foam stabilizers and a gas, usually nitrogen, into the slurry. This creates a low

density slurry that, when properly mixed, forms a very stable and lightweight cement

containing discrete air bubbles that do not coalesce (Crook, 2006).

Corrosion of casing and degradation of cement

Depleted oil and gas wells are favourable sites for CO2 capture and storage (CCS). This process

involves the injection of large amounts of the carbon dioxide, produced from the burning of

fossil fuels, into the Earth. The other common method of CCS is to inject the CO2 into deep

saline aquifers (Carbon Capture & Storage Association, 2015). Over longer timescales (i.e.,

years), a concern the companies responsible for carrying out carbon storage have is the

potential for loss of barrier and well integrity. If a loss of boundary cement integrity occurs

and CO2 is allowed to come into contact with the casing, corrosion can take place (Choi et al.,

2013). If sufficient interaction takes place, complete well integrity failure can occur, resulting

in CO2 leaking into the surrounding rock formations (Choi et al., 2013). It should be kept in

mind that the information presented in the remainder of this section has been obtained from

academic publications related to CCS and not shale gas extraction. Therefore, caution must

be used when applying the results to shale gas operations. However, these studies are useful

in giving an indication of the potential risks.

Carey et al. (2010) carried out an experimental study on corrosion along the interface

between steel well casing and cement in a synthetic system with a fluid composed of 50:50

CO2 and NaCl-rich brine. They found that both the cement and casing were affected and that

the fluid had penetrated as far as 250 µm into the cement. The casing showed clear signs of

corrosion that penetrated 25-30 µm into the steel. Precipitates of both calcium and iron were

found on the casing surface, these were considered to represent the products of corrosion.

The rate of corrosion front propagation was calculated to be between 0.4-1.0 mm/yr.

Talabani et al. (2000) discussed the four main types of casing corrosion that can take place in

wells. The first was erosion corrosion which most commonly occurs during, or shortly after,

drilling has taken place and the casing has been inserted in to the borehole. Wear and

abrasion of the casing occurs when salts and oxides are formed or deposited on the surface

of the casing. The risk of casing corrosion taking place can be minimised by carrying out

cementing as soon as possible after the casing has been introduced into the borehole. The

second way corrosion can take place is through evolved hydrogen diffusion into the casing

resulting in embrittlement, and subsequent failure. The third mechanism of corrosion is

through localised growth of bacteria colonies on the casing as these can alter the chemistry

of the casing surface. For example, some bacteria are able to reduce sulphate, producing H2S

38

which can accelerate the rate at which steel corrosion takes place (Sherar et al., 2011). In the

upper areas of the well where multiple layers of casing are present, galvanic corrosion can

take place if the annulus is not properly cemented and a gap forms between two layers of

casing (Figure 12, bottom left). If this annular gap contains an electrolyte (e.g., Na+, Ca2+,

Mg2+) then the casings will effectively connect electrically with one of the casings acting as an

anode (Choi et al., 2013). The overall result is enhanced corrosion. Choi et al. (2013) noted

that if this type of corrosion does take place, it cannot easily be repaired.

Choi et al. (2013) also described ways in which cement can effectively repair itself over time.

For instance, stress cracks in the cement create pathways for fluids to interact with casing

(Figure 12, bottom right). Upon injection of CO2 into the well, reactions can take place with

the cement to form CaCO3 which can fill the cracks. However, this appears to be exclusive to

CCS and the more complex nature of the fluids used during hydraulic fracturing mean that

this type of self-healing is highly unlikely to take place.

Well integrity testing

Detection of barrier or well integrity failure can be carried out by testing for pressure

retention within the casing, known as “sustained casing pressure” at the ground surface

(Davies et al., 2014). This involves increasing the internal pressure of the well to the

operational pressure and observing for any loss of pressure (The Royal Society and the Royal

Academy of Engineering, 2012). A drop in pressure indicates that well integrity has been

compromised. However, this method does not indicate whether complete well failure has

taken place or at what depth failure has taken place. Failure of the cement, leading to the

formation of channels, can be detected through a number of tests. Any fluids that leak from

the well and migrate up through the annulus between the cement and wall rock can be

detected by inserting probes into the soil directly adjacent to the well. Sampling nearby

groundwater can also reveal any leaks (Davies et al., 2014).

Formation Integrity Tests (FIT) can also give an indication of well integrity. This test was

undertaken at the Preese Hall site. It is carried out after the casing has been cemented in

place but before the next section is fully drilled. The operator drills down through the cement

plug at the base of the previous section, drilling continues into the underlying formation but

only for a few tens of meters and the drill is withdrawn. Similarly to the sustained casing

pressure test, FIT involves closing the blowout protector and pumping drilling fluid down the

wellbore at progressively higher pressures, up to the maximum pressure used during drilling.

This, in turn, applies hydraulic pressure to the newly drilled rock and the cement at the

bottom of the previous hole section (Cuadrilla Resources Ltd, 2012). FIT tests give an

indication of any leaks, and hence any problems with the cement or casing installation, while

also providing evidence as to whether the next formation will allow fluids to flow into it. The

FIT is an industry standard test that allows the determination of what is known as the

“equivalent mud weight” (i.e., the optimal average mud weight (in pounds per gallon (ppg) to

39

be used in the drilling fluid) required during the next drilling stage to prevent the collapse of

the borehole due to the pressure of the surrounding rock and to control the reservoir pressure

and prevent in influx of hydrocarbons (Cuadrilla Resources Ltd, 2012).

A cement bond log (CBL) is the main way of testing the quality of the cement job in the

production section and allows identification of any areas that require remedial cementing.

This test was carried out at Preese Hall. As long as properly carried out, CBL tests can give

highly reliable estimates of well integrity and zonal isolation (PetroWiki, 2014a). The tests

work on the principle that acoustic wave amplitude is rapidly attenuated in good cement, but

not attenuated when the cement is poor, i.e., partially bonded (PetroWiki, 2014a). The

quality of a cement job can be divided into one of four scenarios (PetroWiki, 2014a);

1. Free pipe - No cement between casing and cement.

2. Good bond - Cement is properly bonded to the casing and rock formation.

3. Bond to casing only - Cement is bonded to the casing but not the formation.

4. Partial bond. - Space exists within an otherwise well-bonded casing

The CBL test involves lowering a piece of equipment, consisting of sonic transmitters and

receivers, down the wellbore and measuring how well the cement has bonded to the casing

and the wall rock. This is done by transmitting sonic waves through the cement and casing of

the well. Any gaps in the cement will be detected and appear on the resulting log. Cuadrilla

point out that the equipment is highly sensitive and can detect gaps as little as 0.05 mm wide

that are too thin to transmit fluid. They say that even if the cement mix is perfect and

complete filling of the annulus takes place (as shown by cement returning to the surface) the

CBL will still pick up these anomalies (Cuadrilla Resources Ltd, 2012).

There is a number of limitations with conventional CBL techniques. These are outlined in

Table 6. A more modern technique that has been developed is that of radial-cement

evaluation which was developed in order to overcome some of the limitations of conventional

CBL techniques (PetroWiki, 2014a). These devices use at least one azimuthally sensitive

transducer to evaluate the quality of a cement job around the circumference of the casing.

The devices can also use ultrasonic tools to evaluate the cement job by determining the

acoustic impedance of the material outside the casing; if the cement job is good and the

cement is well bonded to the casing, the impedance is high; if the job is poor and there is poor

bonding, the impedance is low (PetroWiki, 2014a). Utilising this ultrasonic technique avoids

the problem of surface roughness and residual cement on the interior of the casing. A major

advantage of this technique is that azimuth “maps” of the cement quality around the

circumference of a well can be generated (PetroWiki, 2014a). Despite these advantages,

radial-cement evaluation is still sensitive to gas that may be trapped in any microannuluses

that form between the cement and casing (PetroWiki, 2014a). This is a problem that is also

faced by conventional CBL techniques.

40

Table 6. Table outlining the limitations of conventional CBL techniques. (PetroWiki, 2014a).

A number of additional tests are carried out to ensure that the well is functioning correctly

(e.g., a blowout preventer test) and that it can withstand the maximum differential pressure

experienced by the well (positive and negative pressure tests). The 2012 UK Oil and Gas

(UKOG) (Oil and Gas UK, 2012) well integrity guidelines say that the testing procedure should

include the success / failure criteria and the reaction to trends, e.g., an increase in annular

pressure.

Pressure testing should be carried out with liquid where possible. Preferably, this should be

water as the use of drilling mud or other fluids containing solids can hide the presence of

small leaks.

The blowout preventer test (also known as a seal assembly test) is carried out to test the

integrity of the interface between the casing and the wellhead (National Commission on the

BP Deepwater Horizon Oil Spill and Offshore Drilling, 2011). To carry out the test, a packer is

installed below the seal assembly which isolates the area around the top of the well and the

seal assembly. The blowout preventer is sealed and the pressure within the well is raised.

The well passes the blowout preventer test if the pressure within the isolated part of the well

remains approximately constant (National Commission on the BP Deepwater Horizon Oil Spill

and Offshore Drilling, 2011). A drop in pressure indicates that there is a leak, and therefore

that the well fails the test.

A positive pressure test involves sealing off the production casing by closing valves in the

blowout preventer. This is designed to test the well’s ability to hold pressure. The pressure

is initially raised to between 200-300 psi within the production casing at which point the

pumps are stopped and the pressure held constant (National Commission on the BP

Deepwater Horizon Oil Spill and Offshore Drilling, 2011). This is known as a low pressure test.

These are important because low pressures are usually a more severe test for equipment

designed to operate at high pressures; failure at low pressures should not pose a safety

hazard; and the barrier has to seal at low pressures (Oil and Gas UK, 2012). If the pressure

remains constant, the well is considered to have passed the low pressure test. The pumps

are then restarted and the pressure increased to the full test pressure and held for a second

41

time (Oil and Gas UK, 2012). As with the low pressure test, if the pressure holds, the well

passes the positive pressure test.

Negative pressure tests (also known as inflow or drawdown tests) are essentially the inverse

of a positive pressure test (National Commission on the BP Deepwater Horizon Oil Spill and

Offshore Drilling, 2011). During the test, the pressure within the well is reduced to a level

below that of the surrounding rock formation (National Commission on the BP Deepwater

Horizon Oil Spill and Offshore Drilling, 2011). This is intended to test whether hydrocarbons

from the target rock formation will flow into the well. The test is failed if the pressure within

the well increases when the well is sealed, or if flow from the well takes place when the well

is open (National Commission on the BP Deepwater Horizon Oil Spill and Offshore Drilling,

2011). In a successful negative pressure test, there should be no pressure increase and no

flow from the well, i.e., the pressure remains approximately constant within the well

(National Commission on the BP Deepwater Horizon Oil Spill and Offshore Drilling, 2011). This

test determines the quality of the wellhead assembly, casing, mechanical seals and cement

seals; it is the only pressure test that evaluates the integrity of the primary cement seals

(National Commission on the BP Deepwater Horizon Oil Spill and Offshore Drilling, 2011).

Geochemistry can also be used to monitor well integrity. As demonstrated previously, a

potential mechanism for monitoring the integrity of hydraulic fracturing wells is through the

use of isotopic measurements of radioactive tracers in flowback fluid and groundwater.

Warner et al. (2014) hypothesised that injection of hydraulic fracturing fluid into shale results

in specific radioactive isotopes entering the fluid. If well integrity is compromised and a leak

occurs, the concentration of these isotopes in ground water should increase.

To investigate this, the authors aimed to establish the chemical and isotopic composition of

both the flowback fluid and the groundwater. By doing this, one can establish whether

flowback fluids have a unique geochemical fingerprint. This, in turn, allows leaks in the well

casing to be identified. Warner et al. (2014) examined water samples from shale gas wells

drilled in the Marcellus Shale and from the surrounding Appalachian basin water supply. They

found that, although the geochemical signature of the flowback fluid is similar to that of the

water produced from other oil and gas wells, the flowback had distinct lithium (δ7Li) and

boron (δ11B) geochemical signatures compared to the ratios (Li/Cl and B/Cl). The ratios were

found to be consistent across the Marcellus Shale. The reason for these two elements in

particular being of higher concentration is thought to be due to either the interaction of

fracturing fluid with clays in the shale, brine trapped within the shale and boron associated

with organic matter. The flowback fluid was characterised by higher B/Cl and Li/Cl ratios

together with lower δ7Li and δ11B values. The current literature search has not revealed any

evidence of this specific technique being applied elsewhere. A problem with this technique

is that it relies on the analysis of groundwater to determine whether a leak has taken place.

Thus, once a leak has been detected, contamination would have already taken place.

42

Evidence of well integrity failure in the UK

There is no evidence of well integrity failure associated with onshore shale gas wells in the

UK, perhaps due to the limited amount of drilling that has taken place. As a result, evidence

of onshore well integrity comes from conventional hydrocarbon wells. Between 2000 and

2013, the EA recorded nine onshore incidents of pollution associated with the release of crude

oil within 1 mile of the well in question (Davies et al., 2014). Of these nine incidents, two

were associated with well integrity problems, both of which took place at the Singleton Oil

Field in 1993. The current operator by IGas but the field was under different ownership when

the incidents occurred. The failure was detected via groundwater monitoring. A failure

adequately to cement the conductor and intermediate casings (Figure 8) led to a leak, the

magnitude of which was not stated. In terms of environmental impact, no air or land

contamination occurred and minor water contamination was found (Davies et al., 2014). An

investigation, overseen by the Environment Agency (EA), began in 1997 and the source of the

leak was identified. Remedial action was taken and d monitoring has been continued at the

site. Currently, levels of contamination are within the acceptable range set by the EA (Davies

et al., 2014). The other leaks were caused by problems with the pipeline linked to the well.

No leaks were reported from abandoned sites. However, a caveat to this is that abandoned

wells in the UK are currently not monitored and pollutants that are less obviously visible, for

instance colourless and odourless methane, are less likely to be reported (Davies et al., 2014).

It is evident that this lack of monitoring will have to change if shale gas exploration goes ahead

in the UK.

Evidence of well integrity failure in the US

Estimates from the Marcellus shale in Pennsylvania suggests that barrier integrity failure took

place in 3.4% of wells (219 of 6466) drilled between 2008 and 2013 (Vidic et al., 2013). Of the

4602 wells drilled between 2010 and 2012, 7% (320) displayed a loss of integrity (Ingraffea,

2012). However, the latter author gave no indication of how many of these wells that

experienced barrier integrity failure required remedial action or constituted contamination

incidents. This was consistently higher than the failure rate of conventional wells drilled

during the same time period. The cause of failure was deemed to be a combination of poorly

installed, insufficient and defective cement and / or casing. Ingraffea (2012) highlighted the

fact that well integrity was not improving with time, suggesting that the US operators were

not adhering to best operating practices or regulations were not being successfully enforced.

Ingraffea et al. (2014) conducted an in-depth review of 75,505 compliance reports for 41,381

conventional and unconventional oil and gas wells drilled in Pennsylvania, USA, between

January 2000 and January 2013 with the aim of investigating failures in well integrity through

casing and cement problems. Wells associated with shale gas exploration and extraction

43

exhibited a six fold increase in the incidence of cement and / or casing problems when

compared to problems associated with conventional wells.

Wells drilled between 2000 and 2012 associated with both conventional and unconventional

oil and gas wells displayed a rate of cement and / or casing failure of 1.9% across the entirety

of Pennsylvania. Conventional wells displayed a 1.0% failure rate while 6.2% of

unconventional wells lost integrity over the same timeframe. However, upon closer

examination, it is apparent that there are clear geographical differences in the failure rates.

For instance, the majority of unconventional wells are located in the north-eastern area of

the state. Wells located here had a much higher failure rate; 5.21% of conventional wells

drilled between 2000 and 2009 failed, while after 2009, 2.27% of wells lost integrity compared

to 0.73% and 2.08% of wells drilled in the remainder of the state over the same timeframes.

The same was also the case for unconventional wells. Wells drilled prior to 2009 in north-

eastern counties exhibited a failure rate of 9.84% which dropped to 9.14% for wells drilled

after 2009. This is a substantial increase over the rest of the state where failure rates of 1.49%

and 1.88% are observed for unconventional wells drilled pre-2009 and post-2009 respectively

(Ingraffea et al., 2014). The authors suggest that the high occurrence of well failures in

unconventional wells in the northeast of the state between 2000 and 2009 may be attributed

to the fact that the industry was inexperienced in drilling the Marcellus Shale. In addition,

only 61 unconventional gas wells were drilled during this period. Therefore, the failure rates

may be unreliable on account of the small sample size. However, the small reduction in failure

rates in the 2,714 unconventional wells drilled since 2009 suggest that this is not the case

(Ingraffea et al., 2014).

The authors attributed the increased failure rate in post-2009 unconventional wells to

increased pressure on the operator to start production as soon as possible, resulting in the

installation process, and consequently well integrity, being compromised. However, the

authors point out that the increased failure rate could also be due to human factors such as

increased awareness of the issue of well integrity leading to well inspectors carrying out a

more thorough job (Figure 13) (Ingraffea et al., 2014). Indeed, in the US, the percentage of

wells that are inspected within the first year of operation increased from 76% prior to 2009

to 88.7% after 2009 (Ingraffea et al., 2014). It also has to be considered that if a well is found

to have a leak; future inspections may be more thorough, resulting in an increased chance to

find leaks, particularly smaller ones that may have been missed by less thorough inspections.

With this in mind, the authors warn against direct comparison between failure rates between

old and new wells without these factors being considered.

44

Figure 13. Chart displaying the number of wells inspected per year and the percentage of wells constructed per year in which an indication of a failure in well integrity was found. (From Ingraffea et al., 2014).

The integrity of wells at underground gas storage sites in the US has been the subject of a

review by Miyazaki (2009). These sites usually take the form of depleted oil and gas fields,

mining caverns and aquifers, and are used to store surplus gas. Several documented instances

of gas leaking from these storage facilities are detailed by Miyazaki (2009). For instance, the

Magnolia facility near Grand Bayou, Louisiana, where a loss of casing integrity at 442 m depth

resulted in a gas leak and the evacuation of 20 homes approximately 2 miles from the facility.

The leaks documented at CCS facilities were attributed to either ruptured well casings,

separation of the well casing or damaged seals. These factors can influence well integrity on

a long term basis. However, there are factors that can influence integrity on a shorter term

timescale. For instance, in terms of the local geology, the presence of high pressure

formations or aquifers at shallow to intermediate depths can influence the integrity of the

annular seals (Miyazaki, 2009).

Miyazaki (2009) point out that most wells, even if abandoned according to government

standards, will develop leaks with time. This is understandable and expected, to a degree if

the well was constructed in the mid 1900’s, as is the case for many wells in the US (Miyazaki,

2009). However, it has been documented that approximately 10% of abandoned wells in

California develop leaks within 10 years of abandonment (Miyazaki, 2009). However, the

author does not quantify the severity of the leaks or what, if any, remedial action was taken.

Low pressure leaks are most often monitored and periodically vented at the ground surface

when the pressure reaches an appropriate level (Ingraffea et al., 2014). However, if the

process of venting is carried out until the well pressure is zero, the process can lead to

migration of gas into the annulus (Kinik and Wojtanowicz, 2011). If the well pressure

45

increases directly after the well has been vented, the leak is considered to be high pressure

and remedial action should be taken. This can involve re-cementing the problem area (known

as cement squeeze), pumping polymer gels into the problem areas (gel squeeze), sealing the

leak using packers or topping off the cement (Ingraffea et al., 2014).

Current US National Energy Technology Laboratory research

The US Department of Energy (DOE) and National Energy Technology Laboratory (NETL) are

currently funding four projects related to well integrity and zonal isolation, all of which are

due to end between December 2015 and October 2017. The projects are as follows (National

Energy Technology Laboratory, 2015e);

“Nanite” for better well-bore integrity and zonal isolation. Nanite is a cementitious

material containing a functionalised nanomaterial additive that, when mixed into the

cement slurry, can transform cement into a smart material capable of sensing and

responding to stress, pressure and temperature changes in addition to any change in

cement composition. Changes can be detected through electrical, acoustic and

electromagnetic measurements (National Energy Technology Laboratory, 2015c).

Development of nanite will allow information regarding cement barrier integrity,

direct measurements of casing stress, cement shrinkage and well conditions

throughout the life-cycle of the well and identify any infiltration of gas / fluid / wall

rock into the cement.

nXis well integrity inspection in unconventional gas wells. This project involves the

development of a combined X-ray / neutron backscatter imaging device capable of

providing information about well and barrier integrity that is of higher accuracy than

the currently available techniques (National Energy Technology Laboratory, 2015d).

MCIP - this project is focused on improving the current methods of remediating

leaking wells through the use of microbially-induced calcite precipitation (MICP).

MICP has been shown previously to be capable of sealing a downhole fracture in

sandstone. However, it has yet to be seen whether this can be applied to poorly

cemented wells (National Energy Technology Laboratory, 2015b).

Annular isolation in shale gas wells: prevention and remediation of sustained casing

pressure and other isolation breaches. This project aims to develop techniques to

improve groundwater contamination mitigation, to improve failed well seal

remediation, to improve annular isolation during well construction and to develop

techniques to avoid annular seal failure (Research Partnership to Secure Energy for

America, 2014).

46

Evidence of water contamination

The issue of groundwater contamination has proved to be a contentious one in the US. For

instance, there have been studies carried out in the areas overlying the Marcellus Shale,

Appalachian Basin, Pennsylvania that examined the potential for contamination of

groundwater by methane (Osborn et al., 2011; Jackson et al., 2013; Molofsky et al., 2013;

Darrah et al., 2014). This contamination has been proposed to be directly linked to shale gas

drilling and hydraulic fracturing (Osborn et al., 2011). However, there are other studies that

have disputed the claims made in these investigations (e.g., Molofsky et al., 2013).

In an extensive review of the published information on the hazards and risks posed to water

resources by shale gas exploration and extraction carried out by Vengosh et al. (2014), four

main areas of hazard were cited;

1. Contamination of shallow aquifers by fugitive methane and saline deep formation

waters which can result in groundwater becoming salinated.

2. Spills or leaks of inadequately treated flowback and wastewater on the surface

and / or leaks from joints in pipes and from storage tanks, this can result in

contamination of shallow groundwater and surface water.

3. Accumulation of either radioactive elements or toxins in sediments and soils near

disposal, leak or spill sites resulting in contamination.

4. Over-extension of the water supply due to the high demand for water during the

hydraulic fracturing stage that can lead to water shortages and conflicts with other

water uses.

A significant challenge with regard to determining the risk of groundwater contamination

taking place involves the issue of groundwater flow rate. Shale gas drilling in the US has only

accelerated dramatically since the early 2000’s, but the rate of groundwater flow is too slow

for any large scale contamination to have yet become apparent (Vengosh et al., 2014).

However, there is evidence of groundwater contamination from conventional oil and gas

extraction that dates back much further, thus providing a potential indication of how

contamination could take place or develop in future (Vengosh et al., 2014).

Water contamination associated with industrial activity

There is evidence for groundwater contamination taking place in other industrial settings

directly linked to energy production. For instance, the process of harnessing geothermal

energy has been shown to result in groundwater contamination in Turkey (e.g., 2000; Dogdu

and Bayari, 2005; Aksoy et al., 2009). Aksoy et al. (2009) attributed the contamination to four

mechanisms;

47

(a) natural upwards movement of geothermal (hot) fluids along faults, driven by

expansion during cooling,

(b) accelerated upwards migration of fluids due to poorly constructed boreholes,

(c) faulty reinjection applications and

(d) uncontrolled discharge of waste products into the local drainage network.

They also cited contamination by heavy metals (e.g., arsenic, antimony and boron) as the

most significant potential contaminant associated with geothermal energy.

Coal mines have also been demonstrated as being a source for groundwater contamination

through acid mine drainage (e.g., Johnson and Hallberg, 2005; Akcil and Koldas, 2006; Bhuiyan

et al., 2010). Acid mine drainage refers to the process of water flowing through active or

abandoned mines and, in the process, leaching both heavy metals from the surrounding rock

formations and acidic slats that have built up in the wall rock (Johnson and Hallberg, 2005).

Waste produced from mining operations can also be a source for acidic, heavy metal enriched

fluids if inappropriately stored or disposed of, i.e., left open to the atmosphere at the Earth’s

surface.

Water contamination associated with shale gas operations

There exists the remote possibility that fractures created during the hydraulic fracturing

process can penetrate into shallow groundwater. The fractures can act as conduits for

fracturing fluids, gas and other contaminants to enter the aquifer. The risk of this taking place

is minimal enough to be considered acceptable as hydraulic fracturing usually takes place

kilometres below the aquifer and the fractures typically extend, mainly laterally, to a few

hundred meters (Davies et al., 2012; Department of Energy and Climate Change, 2014c). In

addition, a pressure gradient is established around the rock formation when hydraulic

fracturing takes place, this means that flow into the well will preferentially take place, not

flow away from the well.

The largest vertical hydrofracture propagation reported from US hydraulic fracturing

operations is 588m (Davies et al., 2012) (see Text Box 1 for a discussion of what controls the

vertical extent of fractures). By analysing data from five US shale formations, the authors

determined that the probability of a fracture propagating vertically more than 350m is

approximately 1%. The data was obtained from previous microseismicity studies. However,

the authors did not have access to the primary data set. This introduces uncertainties in the

presented results. Firstly, measurements of the vertical extent of the fractures were made

directly from the published graphs. This decreases the likelihood of the resolution of the

smaller fractures. Further, the resolution of fractures smaller than 100m is limited owing to

the technical limit on maximum fracture resolution attainable by the microseismic analysis.

The impact of these uncertainties is that there is considerable error in the recognition and

measurements of fractures smaller than 100m. It is likely that the number of smaller fractures

48

is underestimated and that, in fact, the chance of a fracture propagating more than 350m is

likely to be much less than 1%. The authors justify the uncertainties by highlighting that it is

the largest fractures that are of most interest and that there is considerably less error

associated with these fractures as their vertical extent is easier to measure.

Figure 14. Cumulative graph (horizontal scale) displaying the vertical extent (VE) of stimulated hydraulic fractures together with their baseline fracture initiation depth in five major shale gas bearing formations in the USA. (From Davies et al., 2012).

49

Text Box 1 – The vertical extend of fractures

Fractures form, and propagate, in a plane containing the direction of maximum principal stress (Figure 15). Because of the

depth at which hydraulic fracturing takes place, the direction of maximum principal stress is normally vertical; therefore, the

fractures propagate in the vertical plane, at right angles to the least principal stress. (Hubbert and Willis, 1972).

The in situ stress field (described by the magnitudes and orientations of the principal stresses in three orthogonal directions)

at the depth at which hydraulic fracturing takes place not only control the direction in which the fracture propagates, but

also the fracture shape, the vertical extent of the fracture, and the hydraulic pressure required to initiate the fracture.

Because the fracture orientation is determined by the orientation of the principal stresses, if the fracture propagates into a

layer where the maximum principal stress is not vertical, i.e., on a bedding plane or similar structure, the fracture will

re-orientate and continue propagating (Frac Focus, 2015a). Note that for a fracture to form initially, the pressure at which

the fracturing fluid is pumped at will have to exceed the minimum principal stress plus the tensile stress of the target

formation (Nolen-Hoeksema, 2013).

The vertical extent of fractures is determined by the formations either side of the target formation together with the volume,

rate and pressure at which the fracturing fluid is pumped into the well (Frac Focus, 2015a). If the formations either side of

the shale are stronger or sufficiently stiff then the fracture will not propagate further. Indeed, the fracture may propagate

laterally along the contact between the two formations. It has been suggested that the volume of hydraulic fracturing fluid

pumped into the well is the main constraining factor that dictates the distance the fault migrates (Flewelling et al., 2013).

Figure 15. Diagram of the three principal directions of stress and how they influence the orientation of the fracturing formed during the hydraulic fracturing process. The maximum principal stress is vertical in this case (𝝈𝒗) the two horizontal principal stresses are not equal, i.e., there is a maximum (𝝈𝑯𝒎𝒂𝒙) and minimum stress (𝝈𝑯𝒎𝒊𝒏). Because the maximum principal stress is vertical, the fractures propagate

in the vertical plane normal to H min, but fractures grow most in the horizontal direction. . (From Nolen-Hoeksema, 2013).

50

There are examples of extreme events where the prolonged re-injection of waste water into

a well can lead to larger, unintentional fractures that can breach the surface. Note that this

is not hydraulic fracturing, but a similar activity that can be used as an analogue. At the Tordis

Field near Norway, re-injection of 1,115,000 m3 of fluid over the course of 5.5 months has

resulted in 900 m long fractures forming from the re-injection point to the sea floor (Løseth

et al., 2011). However, the amount of water re-injected into the well was more than 120

times greater than that typically used in hydraulic fracturing (Løseth et al., 2011). Currently,

no examples of this degree of fracturing have been observed onshore.

One way to mitigate the risk of fractures intersecting aquifers would be either to minimise,

or stringently regulate, the amount of water re-injected into the well. It should be noted that

even if these large fractures occur, they would have to extend for more than one kilometre

vertically before direct or indirect connection with mobile shallow groundwater takes place.

Indeed, it has yet to be proven whether groundwater contamination can be caused directly

by upward propagation of fractures resulting from hydraulic fracturing (Davies, 2011)

although the possibility of it taking place cannot be completely dismissed (Davies et al., 2014).

There is also concern that hydraulic fracturing could allow upwards migration of fluids and

gas through the rocks resulting in the potential for groundwater contamination. The geology

of the UK usually means that it is likely that there will be at least one impermeable layer of

rock between the fractures and any aquifer (Department of Energy and Climate Change,

2014e). This is the case at Preese Hall, where the well was drilled through the impermeable

Manchester marl (see the Preese Hall Case Study Section). As these impermeable layers lie

above the shale formation, they will have to be drilled through. Therefore, it is imperative

that well construction and completion is of a sufficient standard so as to minimise the risk of

fluid flowing up through the annulus between the rock and the outer well casing.

Contamination by upwards migration of fluids

Upwards migration of liquids from deep formations into shallow aquifers has been cited as a

potential consequence of shale gas operations. It is to be expected that upwards migration

of liquids is unlikely, simply because they are dense (the same may not apply to gas). Warner

et al. (2012) carried out a study examining the evidence for potential upwards migration of

brines originating from deep formations into shallow water aquifers. The authors examined

the geochemistry of a total of 426 groundwater samples from north-eastern Pennsylvania.

83 brine samples from the Appalachian basin were also analysed, in addition to previously

published geochemical information on deep brines, and used for comparison. The authors

tested the samples for δ18O, δ2H, 87Sr/86Sr and 228Ra/226Ra isotopic tracers.

Groundwater samples showed little difference from meteoric (rain) water in terms of δ18O

and δ2H composition (oxygen and hydrogen isotopic deviations from standard ocean water)

(Warner et al., 2012). However, some samples did display strongly increased salinity

51

(Cl- >20 mg/L) which the authors attributed to brine migration along natural conductive

pathways, as there was no correlation between shale gas drilling activity and the location

from where these high salinity samples were taken. Although this study suggests overall that

the contamination of groundwater due to fluid migration from depth following from shale gas

activity is unlikely to take place, it does raise the point that these conductive pathways can

exist, and if brine can migrate upwards along them, then there is a chance that hydraulic

fracturing fluid and methane could do the same. The net result is that in areas in which

upwards migration of deep formation waters can take place there may be a greater risk of

shallow groundwater contamination (Warner et al., 2012). The authors suggest that future

research should include monitoring of areas of known upwards migration in order to establish

the source of the deep fluid, the connectivity of the pathways and the timescale over which

the upwards migration takes place.

The US National Energy Technology Laboratory (NETL) has recently published a study

examining hydraulic fracturing-related fracture growth and gas / fluid migration in Greene

Country, Pennsylvania (Hammack et al., 2014). The study involved monitoring seven wells

drilled into Upper Devonian / Lower Mississippian rocks overlying the fracturing area of six

horizontal Marcellus Shale wells (Figure 16); in addition, two vertical monitoring wells were

drilled into the Marcellus Shale. In order for migration of fracking and formation fluids to take

place, and hence for contamination of the groundwater to occur, the Tully Limestone (which

is considered to be the barrier that prevents fracture and fluid migration), located

approximately 85 m above the Marcellus Shale, would have to be penetrated by the fractures

and fluid (Figure 16) (Hammack et al., 2014).

Monitoring took place during and after hydraulic fracturing. Gas pressure and production

histories of three of the wells drilled into the Upper Devonian / Lower Mississippian rocks

showed no change in the 12 month period after fracturing ceased. This indicates that no

upwards migration of fluid took place, as a change in pressure would indicate communication

between the fluid / gas and the over-pressurised (meaning the gas pressure in the sale is very

high) shale (Hammack et al., 2014). Additional monitoring of possible background migration

of fluids took place two months before the fracturing took place. Measurements of gas and

produced water samples were taken from the seven wells drilled into the Upper

Devonian / Lower Mississippian rocks, as well as from the two vertical Marcellus Shale wells,

throughout the fracturing stages and for eight and five months respectively after fracturing

ended. Gas samples were examined for carbon (δ13CCH14) and hydrogen (δ2HCH4) isotope

signatures and produced water was tested for its strontium isotope composition. Clear

differentiation in terms of isotopic signatures can be seen between the two sets of wells

(Figure 17), thus indicating that no mixing, and hence no upwards migration, took place

(Hammack et al., 2014). The isotopic signatures also show little variation with respect to time

(Figure 17), this is also the case for the strontium isotope ratio of the produced water taken

from the Upper Devonian / Lower Mississippian wells (Figure 18) (Hammack et al., 2014).

52

Four perfluorocarbon tracers were also utilised during the fracturing stage of the operation

in order to establish whether upwards migration of fluids took place. The tracers were mixed

with hydraulic fracturing fluid and injected into a horizontal Marcellus Shale well at 10

fracturing stages. The introduction of the tracers took place after the well had been acidized

but before the proppant had been introduced into the hydraulic fracturing fluids. This

allowed the tracers to migrate to the most distant parts of the fractures. The tracers were

designed to form a non-aqueous, buoyant phase when entering the rock formations, thereby

facilitating easy upwards migration (Hammack et al., 2014). Tracers were found in the gas

samples taken from the Marcellus well into which the tracers were injected. Additionally,

tracers were found in an adjacent well, approximately 230 ft. from the tracer injection well.

The authors consider this to be a result of the perforation zones of each wellbeing in close

proximity to each other. Samples from the Upper Devonian / Lower Mississippian wells

nearest to the tracer injection well displayed no evidence of the presence of tracers during

the monitoring period of two months, therefore, suggesting that no upwards migration of

gas / fluid took place (Hammack et al., 2014).

Overall, the study found no evidence of upwards gas / fluid migration at any point in the

monitoring programme. Therefore, no evidence of groundwater contamination was found

resulting from fluid migration through the rock succession (Hammack et al., 2014). Continual

monthly (produced water) and bimonthly (gas) measurements are currently being made.

53

Figure 16. Diagram displaying the depth relationship between the shallow groundwater aquifers (USDW), the Upper Devonian / Lower Mississippian formations into which seven vertical monitoring wells were drilled (Monitored Interval) and the Marcellus Shale target formation together with the overlying Tully Limestone which is considered to act as the barrier to fluid and fracture migration (Fractured Interval). Six horizontal production wells, together with two vertical monitoring wells, were drilled into this interval. (From Hammack et al., 2014).

54

Figure 17. Top) Comparison between the isotope signatures of the Upper Devonian / Lower Mississippian monitoring wells and the vertical wells drilled directly into the Marcellus Shale. Note the clear differentiation between each group. Middle) Change in carbon isotope signature over the sampling period. UD-1-UD-7 denotes Upper Devonian / Lower Mississippian monitoring wells while MW-1 and MW-2 denote vertical Marcellus Shale wells. Bottom) Change in hydrogen isotope signature over the sampling period. The key for the middle and bottom plots are the same. Note that the levels in both the middle and bottom plot remain approximately

55

constant, Hammack et al. (2014) consider that the variation is within expected range. (From Hammack et al., 2014).

Figure 18. Plot of the change in strontium isotope ratio with time in five Upper Devonian / Lower Mississippian monitoring wells. Note the approximate consistency of the ratios both pre- and post-fracturing. (From Hammack et al., 2014).

A recent study aimed to model the potential for groundwater contamination by hydraulic

fracturing fluids through the use of pressure changes reported during gas well operation

(Myers, 2012). The study found that the changes in pressure associated with the process of

hydraulic fracturing could allow the upwards migration of fluids into aquifers in less than 10

years. However, as is a limitation of most modelling studies, simplifications must be made in

order to allow the model to be run (Vidic et al., 2013). These simplifications, although

necessary, limit the applicability of the results. In a comment on the study, Saiers and Barth

(2012) argued that key hydrological processes were neglected or misrepresented. For

instance, Myers (2012) assumed that the voids in the shale are filled only with water, whereas

in reality, the voids in the Marcellus and overlying formation contain both water and gas. The

result is that hydraulic conductivity could be orders of magnitude lower (Saiers and Barth,

2012; Vidic et al., 2013) as a result of capillary pressure effects. In order to understand fully

the potential for upwards migration of fluids over long timescales more such modelling

studies are required. However, it is likely that hydraulic conductivity will vary between drilling

sites, therefore, the results of such studies may be restricted in their applicability to other

shale gas operations.

56

Contamination by methane

Osborn et al. (2011) examined the methane content of aquifers (Catskill and Lockhaven

formations in Pennsylvania and the Genesee Group in New York) overlying the Marcellus and

Utica Shales in Pennsylvania and New York. 60 private drinking water wells were tested for

dissolved gas concentrations. The depths to which the wells were drilled, and hence where

the samples were taken from, varied between 36 to 190 m.

As a side note, the US Department of Interior has set limits on the permissible methane

content of groundwater. At levels of <10 mg/L of methane, the water is considered safe; at

levels of 10-28 mg/L, monitoring of the water supply is required. If the concentration rises

above 28 mg/L, i.e., the saturation point of methane in water at room temperature and

pressure, then immediate action is required to remediate the problem (Vengosh et al., 2014).

Osborn et al. (2011) found that methane was present in 51 of the 60 wells tested. However,

in areas around drinking water wells where more than one drilling site was located within

1 km the average methane content in drinking water was 19.2 mg CH4/L; the maximum

methane (supersaturated) content recorded was 64 mg CH4/L (Osborn et al., 2011). This

placed the average drinking water methane content within the hazard range set by the US

Office of Interior (10-28 mg CH4/L) and the maximum methane content well above the hazard

level (Osborn et al., 2011). The isotopic signature of ground water methane was found to

match that of a deep thermogenic source, i.e., the Marcellus and Utica Shales. Conversely, in

areas around drinking water wells with no active drilling sites within 1 km, the average

methane content was 1.1 mg CH4/L (Figure 19). In addition, the isotopic signature of this

methane indicated a biogenic or mixed biogenic and thermogenic source (Osborn et al.,

2011).

57

Figure 19. Graph displaying the methane concentration in water samples taken from areas of active gas extraction (filled circles) and areas of non-active gas extraction (triangles) as a function of distance from the well. The range of methane concentration above which the US Department of Interior requires remedial action to be taken is shown by the grey area. Note that the distance to the well is determined using the position of the well on the ground surface and does not account for the direction and lateral extent of drilling. (From Osborn et al., 2011).

Osborn et al. (2011) proposed three potential mechanisms by which groundwater

contamination could have taken place in the study area. First, the upwards migration of deep,

gas-rich brines displaced by the hydraulic fracturing process; second, a loss of well integrity

resulting in a leak; and third, the development of new fractures or the re-activation of older

fractures causing them to increase in size. The latter mechanism can also result in fractures

becoming more interconnected and therefore making it easier for gas to migrate upwards.

With regard to the first mechanism, the authors found no evidence of deep brines in shallow

drinking water; therefore they considered this mechanism not to be responsible. A loss of

well integrity was cited as the most likely cause of the contamination. However, the possibility

of interconnected fracture networks assisting the upwards migration of methane cannot be

ruled out.

Osborn et al. (2011) recommended that long term sampling and monitoring of drinking water

from private wells by both the property owners and industry operators must take place. They

highlighted the need for such monitoring and analysis to take place before any drilling in order

to establish reliable baselines. The authors also considered that more studies are required in

order to better understand the mechanisms that control the upwards migration of methane.

Since the publication of this article, there has been some debate about whether the elevated

methane levels found by Osborn et al. (2011) were a natural occurrence or a result of drilling

58

activity. This is due to the lack of previously determined baseline levels of methane from the

study area and the fact that the areas are known naturally to produce methane seeps of both

biogenic and thermogenic origin (Vidic et al., 2013). Biogenic gas is formed at shallow depths

and low temperatures by an anaerobic bacterial decomposition of sedimentary organic

matter. Thermogenic gas is formed at deeper depths by thermal cracking of sedimentary

organic matter into hydrocarbon liquids and gas, and thermal cracking of oil at high

temperatures into gas and pyrobitumen.

According to Vidic et al. (2013) the methane contents of groundwater found by Osborn et

al. (2011) were similar to those found by the US Geological Survey during a previous water

sampling survey which took place between 1997 and 2011 (Figure 20), suggesting that there

may have been pre-existing seeps. In order to provide an explanation for this, Vidic et al.

(2013) cited the fact that there have been approximately 350,000 oil and gas wells drilled in

Pennsylvania, the locations of approximately 250,000 of which are unknown. Therefore it may

be that these unknown wells are leaking and acting as conduits for upwards migration of

hydrocarbons. This emphasises the importance of baseline studies in the UK if shale gas

exploration and extraction are to go ahead.

59

Figure 20. A) Published values for methane in groundwater and the location of the respective sample sites. 239 sites sampled from 1999-2011 are in New York State, 40 sites sampled in 2005 are in Pennsylvania and 170 sites sampled from 1997-2005 in West Virginia. The lack of data points in Pennsylvania is due to Vidic et al. (2013) only including samples taken in 2005. B) Graph of the sample sites shown in A with arrows indicating the average methane content of the samples analysed by Osborn et al. (2011). (From Vidic et al., 2013).

Vidic et al. (2013) highlighted the fact that, at the time of publication, only one study had

compared the chemistry of groundwater before and after drilling (Boyer et al., 2012). The

60

study compared 48 water samples from drinking water wells over Pennsylvania in 2010 and

2011. It was found that no statistically significant difference existed between methane levels

before and shortly after drilling, nor was there any correlation between methane content of

water and distance from drilling sites (Boyer et al., 2012). It should be noted, however, that

this study was not peer reviewed. In addition, although no short term increases were seen,

because of the lack of studies on the timescales and mechanism of methane migration, it may

be that upwards migration of methane takes place over several years and would therefore

not be detected shortly after drilling had been completed. The study did find that the

methane concentration increased in a single water well shortly after well completion. The

authors attributed this to problems with well integrity resulting in a leak, they also noted that

this occurrence is consistent with the average rate of well integrity failure observed in the US

(approximately 3%) and is, therefore, to be expected (Boyer et al., 2012).

The frequency of methane detection in water wells varies considerably with respect to

geography. For instance, in the northeast of Pennsylvania, methane has been detected in 80-

85% of tested water wells (Molofsky et al., 2011; Osborn et al., 2011) whereas in the

southwest only 24% of tested water wells have had methane detected in the potable water

(Boyer et al., 2012). This could be attributed to the small sample size used by the Osborn et

al. (2011) or it could be that the geology in the northeast of Pennsylvania is more susceptible

to the upwards migration of methane (Warner et al., 2012; Vidic et al., 2013). On a wider

scale, these differences in geology could explain why there has been little in the way of

groundwater contamination problems in other shale gas producing areas in the US, i.e., the

Fayetteville shale in Arkansas (Vidic et al., 2013). Given that methane migration does occur

naturally, the above discussion further highlights the importance of baseline studies in areas

of potential shale gas exploration and extraction in the UK.

A more recent study by Darrah et al. (2014), building upon previous studies by Osborn et al.

(2011) and Jackson et al. (2013), examined water taken from drinking water wells in the

Marcellus and Barnett Shale areas known to have elevated methane levels. 113 samples were

taken from the Marcellus region and 20 from the Barnett region. The authors examined the

occurrence of the noble gas isotopes 4He, 20Ne and 36Ar, hydrocarbon (CH4, C2H6, C3H8 and N2)

and chloride (Cl-) contents of the water together with stable isotope compositions (δ13C-CH4

and δ13C2-CH6) and the elemental and isotropic compositions of the noble gases. The aim of

the study was to address two specific questions. First, is the methane found in drinking water

natural or of anthropogenic origin? Second, if fugitive methane contamination is taking place,

how is it happening?

Darrah et al. (2014) found that methane was present in water samples taken from wells in

the Marcellus Shale located >1 km from drilling sites; the methane was found in association

with elevated levels of natural brine components (i.e., He4 and Cl-). Water samples taken

from wells with nearby drilling sites (<1 km away) were either indistinguishable from those

taken from >1 km away, or displayed methane supersaturation together with low salt

61

concentrations. In the case of the samples with elevated natural brine components, the

authors considered these samples to represent natural migration of deep, gas-rich brines to

shallow depths. The methane-supersaturated samples also displayed high noble gas ratios

(i.e., CH4/36Ar and 4He/20Ne) which appear to be independent of Cl- concentration. This

indicates elevated thermogenic gas concentration.

Similar trends were found in samples taken from the Barnett Shale in 2012. Twelve water

wells were sampled, nine of which displayed noble gas ratios that suggested natural migration

of gas-rich brines. The remaining three had elevated CH4/36Ar and 4He/20Ne values,

suggesting anthropogenic contamination related to shale gas drilling and extraction. In order

to confirm the results, the authors re-sampled the twelve Barnett Shale drinking water wells

in 2013. An additional eight new wells were also sampled (total of 20 wells). The newly

sampled wells showed no indication of anthropogenic contamination. Ten of the original

twelve wells displayed minimal change, the remaining two displayed increased CH4/36Ar and 4He/20Ne values with no change in Cl-, in one case the values increased by an order of

magnitude. These results indicate that thermogenic gas, without accompanying brine,

contaminated these water wells in less than one year (Darrah et al., 2014).

The anthropogenically-contaminated water samples from the Marcellus Shale also displayed

decreased levels of 36Ar, 20Ne and N2 compared to non-contaminated samples. In the Barnett

Shale area, the three anthropogenically-contaminated wells sampled in the first sampling

instance also displayed decreased 36Ar, 20Ne and N2 values. The two wells which displayed

progressive contamination between the two sampling periods in 2012 and 2013 also

displayed decreasing 36Ar, 20Ne and N2 values. It should be noted that these decreases were

only seen in wells where the nearest shale gas site was <1 km away (Darrah et al., 2014). The

reduction in the concentration of these gases requires unique hydrogeological conditions.

Specifically, gas has to be introduced into the near-surface aquifer at a pressure of

>1 atmosphere, i.e., at a rate higher than natural groundwater flow (Darrah et al., 2014). The

authors argued that there is no obvious tectonic or hydraulic mechanism in either area that

would allow this to happen. This suggests that the mechanism of contamination was by well

integrity failure (Darrah et al., 2014).

Other evidence for methane contamination associated with shale gas activities comes in the

form of hydrocarbon molecular compositions (C2H6+/CH4) and stable isotope compositions

(δ13C-CH4). Elevated values for each of these are considered to be indicators of restricted

biogenic methane production (Darrah et al., 2014). Indeed, elevated levels were detected in

the samples considered to have undergone anthropogenic contamination by fugitive gas

(Darrah et al., 2014). In addition, the hydrocarbon composition levels of the gas in the

samples taken from the Marcellus Shale region match those of the gas extracted from the

drilling sites and that found in intermediate depth formations.

Darrah et al. (2014) concluded that samples that contained both methane and natural brine

components were the result of natural upwards migration of gas-rich brines. They

62

hypothesise that the natural upwards migration took place in three stages. First, sufficient

thermogenic methane formed in the shale source rock to allow the formation of a free gas

phase. This permitted the trace gases (e.g., 20Ne and 36Ar) to fractionate into the gas phase.

The degree of fractionation is dependent upon the partition coefficient of the gas in question.

In terms of the gases analysed by Darrah et al. (2014), He and Ne have higher partition

coefficients (i.e., lower solubility in fluids) compared with Ar and CH4, therefore, He and Ne

will preferentially fractionate into the gas phase. The next stage involves buoyant upwards

migration of the He and Ne enriched hydrocarbon phase. As this takes place, fractionation

continues to take place and the gas phase becomes progressively enriched with both Ne and

He as the other less soluble gases (Ar and CH4) re-dissolve into the water-saturated crust. The

overall result of this process is elevated 20Ne/36Ar and 4He/CH4 values in the gas phase. This

is observed in gases found at intermediate depth formations in the Marcellus Shale. The final

stage of migration involves the diffusion and equilibration of gases into shallow aquifers.

The groundwater samples considered to have experienced anthropogenic contamination due

to fugitive gas migration display significantly lower 20Ne/36Ar and 4He/CH4 values, suggesting

that migration took place with little fractionation, i.e., rapid upwards migration. The lack of

fractionation indicates that there was minimal interaction between the gas and the

intermediate formations, suggesting that either a failure in well integrity or an annular leak

was the cause. In order to differentiate between the two potential mechanisms, the

geochemical fingerprint was examined. Gas that originates from intermediate depth

formations, i.e., the gas that would be found in the aquifer if an annular leak had taken place,

has a lower δ13C-CH4 value than production gas. In addition, C2H6+/CH4 values are lower than

in production gas. When Darrah et al. (2014) examined these geochemical fingerprints, they

found three clusters of results in the Marcellus Shale and one cluster in the Barnett Shale

corresponded to gas found at intermediate depth formations. This suggests that poor

cementing likely resulted in an annular leak which, in turn, allow the rapid upwards migration

of gas. Four clusters corresponded to the composition of the Marcellus production gas,

indicating problems with well construction resulting in leaks of production gas in and around

the aquifer. One of these clusters exhibited evidence of significant fractionation during

migration from depth, resulting in drastically increased 20Ne/36Ar and 4He/CH4 values. The

authors attributed this to the “underground mechanical... failure” of a well prior to the

sampling taking place. The authors found no evidence of large-scale upwards migration of

gas caused by hydraulic fracturing.

Vengosh et al. (2014) suggested that the severity of any groundwater contamination could be

increased if oxidisation of fugitive methane takes place via bacterial sulphate reduction

reactions. When these reactions take place, they can act to mobilise other, more potentially

dangerous elements such as arsenic from the aquifer formation. The overall result is a further

decrease in water quality. Research to date has not produced conclusive evidence that this

takes place in areas associated with shale gas operations (Vengosh et al., 2014).

63

Another possible consequence of groundwater contamination by stray gas is the formation

of toxic trihalomethane compounds (Vengosh et al., 2014). These are compounds that form

in the presence of organic matter, in which halogen atoms substitute for the hydrogen atoms

in methane. Currently, there are no published studies that have observed these compounds

at shale gas sites. However, if both halogens and organic matter are present in the aquifer

after contamination has taken place then there is the potential for trihalomethane

compounds to form (Vengosh et al., 2014).

Contamination of surface water

Contamination of surface water by flowback fluid is also a potential hazard associated with

shale gas exploration and extraction. As discussed previously, when the hydraulic fracturing

fluid enters the rock formation of interest, it can leach chemical elements and compounds

from the rock. In addition, as discussed above, elevated salinity levels are also commonly

associated with the fluids produced in shale gas operations, and surface waters may become

more saline if a spill or leak occurs. These risks vary spatially as demonstrated in the US,

where the salinity of flowback fluids range from slightly less than that of seawater (total

dissolved salts = 25,000 mg/L), e.g., the Fayetteville Shale (25,000 mg/L), to well above

seawater, e.g., Marcellus Shale (180,000 mg/L) (Vengosh et al., 2014), note that the salinity

of seawater is 35,000 mg/L. It has also been shown that the chloride salinity and levels of

potentially dangerous elements, such as strontium and barium, vary naturally within rock

formations (Chapman et al., 2012; Barbot et al., 2013).

Vengosh et al. (2014) describe three main methods by which surface water contamination

can take place:

1. General leaks and spills of flowback fluid from on-site equipment, such as storage

tanks, pipework and open ponds.

2. Direct, unauthorised discharge of untreated flowback fluid into streams and other

surface water sources.

3. Inadequate treatment of flowback fluid, either on-site or in treatment plants, followed

by discharge into surface water sources.

The occurrence of surface water contamination events has been shown to increase in areas

of high density (>0.5 wells per km2) shale gas operations (Vengosh et al., 2014). Vengosh et

al. (2014) say that the risk of surface water contamination is increased in the instance of shale

gas operations, as gas production can rapidly decrease by as much as 85% within the first

three years of extraction. The result of this is that a sufficient number of wells needs to be

drilled in order to keep production at the required economic level. The more wells that are

drilled and fractured, the more risk there will be of a spill or leak taking place. Wells can also

64

be re-fractured in order to re-stimulate gas flow. This will involve more water being

transported on-site, increasing the risk of a spill or leak taking place.

Around conventional wells in Garfield County, Colorado, the percentage of private drinking

wells that had chloride concentrations higher than 250 mg/L increased from 4% in 2002 to 8%

in 2005 (Vengosh et al., 2014). These increases were associated with increasing thermogenic

methane (methane formed under high pressure and high temperature conditions, i.e., >1 km

below the Earth’s surface) in the groundwater, which, in turn, was associated with an increase

in the number of oil and gas wells in the county (Vengosh et al., 2014). The cause of the

increased salinity and methane content of the groundwater was attributed to leaks, either

from the well or from surface water storage pits (Vengosh et al., 2014). The authors say that,

if contamination were to take place at a shale gas site, the water chemistry of the

contaminated water source should be similar to that of the water in the target shale

formation or the water in intermediate depth formations. However, the authors also say that

there is currently no evidence of increased salinity of groundwater in shale gas sites in north-

eastern Pennsylvania that have experienced groundwater contamination via the infiltration

of stray gas. Hence salinity increase alone does not necessarily imply contamination from

shale gas production. More studies are therefore required in order to establish clearly the

extent of correlation between groundwater salinity and shale gas extraction.

There is documented evidence in the US of illegal disposal of untreated and inadequately

treated flowback fluid being released into surface waters, e.g. in the Acorn Fork Creek in

Kentucky (Papoulias and Velasco, 2013). The release of the fluid was considered to have

caused the death of local aquatic species, in addition to adversely affecting their general

health (Papoulias and Velasco, 2013). A study using the controlled release of hydraulic

fracturing fluids into a forest stream revealed that vegetation near to the stream was severely

affected by the contamination (Adams, 2011). In less than 10 days, vegetation had been

damaged or killed. Over the course of two years, over half of the trees within a 2 km2 area

around the stream were dead and the soil had been contaminated by sodium and chloride,

with the levels being 50-fold higher than before contamination (Adams, 2011). It should be

noted that this sort of disposal is against the law in the UK. Therefore, as long as the law is

followed by the operator, the risk of surface water contamination through deliberate release

should be minimal.

Inadequate treatment of flowback fluid before being discharged into surface waters can result

in contamination. It has been demonstrated by Warner et al. (2013) that treated flowback

water discharged from water treatment sites in Pennsylvania still contained elevated salinity,

naturally occurring radioactive materials (NORM), toxic metals and volatile organic

compounds (VOCs), such as benzene, after treatment. VOCs are organic compounds that

have a high vapour pressure, and therefore low boiling point. They have been cited as a

potential human health risk arising from shale gas exploration and extraction (Bunch et al.,

2014). VOCs are found within many everyday products including paints, disinfectants,

65

cosmetics, glues and pesticides (United States Environmental Protection Agency, 2012). A

report by the New York State Department of Health (2014) found that VOCs can increase the

odour in the local atmosphere; they can also increase the chance of respiratory problems.

VOC emissions can be managed using green completions in the same way as fugitive methane

emissions. In terms of salinity, the concentration of chloride and bromide in the treated fluid

were 80,542 mg/L and 644 mg/L respectively at the point of discharge compared to

background levels (18 mg/L and 0.05 mg/L respectively) present upstream (Warner et al.,

2013). In terms of salinity, the flowback fluid was 2.3 times the salinity of seawater. As far as

2 km downstream from the discharge point, levels of bromide were found to be 16-fold higher

than background levels (Warner et al., 2013).

Bromide poses a potentially significant hazard as its presence can result in the formation of

carcinogenic trihalomethanes in chlorinated drinking water (e.g., Chowdhury et al., 2010a, b).

Evidence from tests on drinking water in some areas of Pennsylvania, specifically the

Monongahela River and Pittsburgh, have shown elevated levels of bromide related

trihalomethanes related to the improper treatment of waste fluids (States et al., 2013; Wilson

and Van Briesen, 2013). If the flowback fluid is transported to off-site facilities for treatment

then it is possible that this lack of appropriate treatment is not the fault of the operator, but

more the fault of either the authorities that decide which water treatment facilities are able

to treat properly the fluid, or the companies who operate the facilities. If shale gas

exploration and extraction goes ahead in the UK, the inspection of the procedures in

operation at these water treatment plants will be important in order to minimise the surface

water contamination hazard. Vengosh et al. (2014) point out that bromide related

trihalomethane can also form in wastewaters associated with other forms of industrial

activity, such as conventional oil and gas operations. On account of this, they recommend

continuing research in order assess to more fully the potential for surface water

contamination from waste waters arising from shale gas operations.

Vengosh et al. (2014) point out that “over time, metals, salts, and organics may build up in

sediments, scales, and soil near wastewater disposal and/or spill sites. Each respective

compound has a given solubility and reactivity (e.g., adsorption), the latter commonly

described by the distribution coefficient (Kd) that varies as a function of pH, Eh, temperature,

and the occurrence of other components in the water. As a result, the physicochemical

conditions of surface waters and the distribution coefficients of each compound will

determine how it interacts with particulate matter (e.g., colloidal particles) or river sediments.

Ultimately, these properties will determine the long-term environmental fate of such reactive

contaminants; reactive constituents would be adsorbed onto soil, stream, or pond sediments

and potentially pose long-term environmental and health risks.”

There is evidence for contamination of sediments located downstream of water treatment

plants in Pennsylvania (Warner et al., 2013). The NORM levels, specifically radium isotopes

(228Ra/226Ra), in these sediments matched those of brines originating from the Marcellus

66

Shale, therefore indicating that there is a direct relation between contamination and shale

gas waste water (Warner et al., 2013). In one instance, the levels of NORM in sediments at

one treatment facility exceeded safe levels for disposal of radioactive material (Warner et al.,

2013). This is potentially significant as the fluid treated at this facility was likely dealt with

using the appropriate procedures, yet elevated levels of contaminants still built up over time.

The UK should be aware of this if shale gas exploration and extraction proceeds. Techniques

for long term monitoring of discharge locations should also be considered.

Groundwater contamination in the UK

Past instances of groundwater contamination in the UK have come from abandoned coal

mines (Younger et al., 2002). Although now-abandoned coal mines were continually drained

whilst in operation, after decommissioning, drainage was stopped. After 10-20 years,

groundwater rebound can take place, resulting in the aquifer becoming contaminated by

saline fluids of deeper origin than could have become acidified by reaction with pyrite, and

can carry large amounts of colloidal iron oxide (the acid-mine-drainage problem). This series

of events took place in the Durham Coalfield where a public water supply was contaminated,

and consequently failed to meet EU groundwater quality standards (Neymeyer et al., 2007).

Abandoned coal mines can also result in the release of methane; in 2008 UK coal mines were

estimated to have produced approximately 14 million m3 of methane (Davies et al., 2014).

In order to minimise the risk of groundwater contamination, environmental regulations have

been put in place that control the design, construction, operation and monitoring of the well

and associated activities. Part of the planning application process is the submission of a

hydrogeological report to the EA. The report details the presence of any ground or surface

water as well as the measures proposed to mitigate any risks to the environment from all

stages of shale gas operations. This requires inclusion of details about the methods used to

construct the well, and the monitoring processes in place to monitor the well integrity.

Additionally, the report must state the composition of the hydraulic fracturing fluid itself, the

presence of any NORM in the local geology, how the water needed for the fracturing fluid is

to be acquired and how the waste from the site will be managed (Department of Energy and

Climate Change, 2014e).

As has been discussed previously, there have been examples of well integrity failure in the US

which has resulted in groundwater contamination (Darrah et al., 2014). It should be noted

that cases of groundwater containing elevated levels of methane have been attributed to

failures in the construction of the well, not to the propagation of a fractures that introduce

pathways to the shallow overlying aquifer (Darrah et al., 2014; Vengosh et al., 2014). It is

considered that if best practices are followed in the UK, the risks of such well-failures should

be minimal. Best practices are outlined by the United Kingdom Onshore Oil and Gas

Operators Group (UKOOG) and DECC (Department of Energy and Climate Change, 2013;

United Kingdom Onshore Operators Group, 2015).

67

The BGS has documented the spatial relationship between UK aquifers and the underlying

prospective shales (Table 7). In addition, the BGS has created a set of 25 maps that display

the vertical separation between the aquifers (whether or not exploited for potable water) and

shales, and how this separation varies laterally. Examples of these maps are shown in Figure

21 and Figure 22 which display the vertical separation between the Bowland Shale and the

Triassic Sandstone aquifer and the separation between the Kimmeridge and Ampthill Clay

formations and the Lower Greensand aquifer formations respectively. In addition to these

maps, the BGS provides information regarding the lithologies of the intervening strata. This

is useful when considering the statement made by DECC (Department of Energy and Climate

Change, 2014e) that there are layers of rock between the aquifers and the shales that are of

low permeability and will thus act as a barrier to any migrating fluids / gas. In terms of Figure

21, the intervening strata to the West of the Pennines consists of the Millstone Grit and Coal

Measures (mudstone, siltstone and sandstone) of mixed permeability, the permeable

Permian Sandstone (another principal aquifer) and the low permeability Permian Mudstones

of the Cumbrian Coast Group. The Millstone Grit and Coal Measures occur as part of the

intervening strata across all areas shown in Figure 21. The Wealden deposits, which are

described as being of low permeability, occur in most of the intervening strata between the

Lower Greensand aquifer and the Kimmeridge and Ampthill Clay (Figure 22).

The BGS are continuing work on this topic to characterise better the

geological / hydrogeological properties of the intervening layers. The objective is to develop

a groundwater vulnerability model that is applicable to activities (such as shale gas) that take

place below aquifers.

68

Table 7. Table the spatial relationship between aquifers and potential shale gas sources. (From Ward et al., 2014).

69

Figure 21. Map displaying the lateral variation in intervening strata thickness between the Triassic Sandstone aquifer and the Bowland shale. (From British Geological Survey, 2015b).

70

Figure 22. Map displaying the lateral variation in intervening strata thickness between the Lower Greensands aquifer and the Kimmeridge and Ampthill Clays. (From British Geological Survey, 2015a)

71

Well abandonment

Many of the same concerns outlined in the well integrity section also apply to well

abandonment. The issue of fluid leakage and fugitive gas emissions will likely become more

important during, and in the time after, well abandonment. During exploration and

extraction, constant monitoring of the well takes place. However, after the well has been

abandoned, it is likely that the frequency with which the well is monitored will decrease. As

time goes on, the risk to the environment potentially increases as cement can undergo

degradation; the casing can also fracture and corrode at the connections between sections

(Davies et al., 2014; Jackson, 2014). Degradation of the cement can result increased

permeability, thus introducing a potential pathway for fluid flow (Davies et al., 2014).

Once all operations at the site have ceased, it is the responsibility of the operator to ensure

that the site is restored to a state similar to that prior to the drilling, or to a condition that

would allow it to be re-visited at a later date (Department of Energy and Climate Change,

2014b). In order for the well to be suitably abandoned, the well must be securely sealed so

that there can be no leakage from within the well bore. In order to seal the well, cement is

pumped into the production casing and a steel cap is fitted to the top of the well (Davies et

al., 2014). A typical abandoned conventional oil / gas well can be seen in Figure 23. Note that

multiple cement plugs are used, one at the depth at which perforation takes place, i.e., the

production zone, and one at aquifer depth. The top of the production zone plug extends into

the caprock formation. This is done in order to ensure complete isolation between the

production zone and the well (Choi et al., 2013).

72

Figure 23. Schematic diagram of an abandoned conventional oil and gas well. (From Choi et al., 2013).

UKOOG has published industry guidelines related to all stages of well development, appraisal

and abandonment (United Kingdom Onshore Operators Group, 2015). UKOOG say that

operators should conform to the Oil and Gas UK Suspension and Abandonment Guidelines.

They also suggest that operators should consider the following factors when planning well

abandonment (United Kingdom Onshore Operators Group, 2015):

Height of the cement in annulus outside casing

Any permeable zones that should be cemented

Areas of cementing casing overlaps

Cement abandonment plugs need to cover the entire diameter of the well. Down-

hole equipment, e.g., cables, should be removed from the well

The type of fluid in the annuli above the cement.

The process of injecting cement into the annulus is more difficult than primary

cementing.

The process of well decommissioning and abandonment has been outlined by the EA

(Environment Agency, 2012a). The EA says that, in order for a well or borehole to be

abandoned in a satisfactory manner, it must be made safe, structurally sound and backfilled.

73

All these measures are to prevent the unplanned release of hydrocarbon and the upwards

migration of fluids into shallow groundwater areas (Environment Agency, 2012a, b). The EA

also differentiates between well abandonment and well suspension. Suspended wells are

those that are temporarily abandoned during parts of the well lifecycle while abandoned wells

cannot be used again (Environment Agency, 2012b). If a well is suspended, it must be

maintained in such a state that a routine work over could restore it to regular operation

(Environment Agency, 2012b).

When the decision to abandon a well is made, both DECC and the Health and Safety Executive

(HSE) have to be notified. The operator must submit an abandonment application to DECC

which must be approved before any work can proceed. This contains information regarding

the specifications of the abandoned well together with an abandonment plan (Environment

Agency, 2012b). Once approved, the HSE must be given at least 21 days’ notice before

abandonment begins (Environment Agency, 2012b).

OGUK, the trade body that represents the UK offshore hydrocarbon industry, published

guidelines on well integrity throughout the lifecycle of extraction operations in 2009 (Oil and

Gas UK, 2009). The EA uses these guidelines, in conjunction with the American Petroleum

Institute’s (API) recommended practice guidelines on environmental protection during

onshore operations (American Petroleum Institute, 2009), to inform the approach that should

be taken with UK onshore shale gas operations. However, the EA’s review of casing

installation at shale gas sites (Environment Agency, 2012b) summarises the relevant

guidelines for well abandonment made in both the OGUK and API reports. These are

summarized below (Environment Agency, 2012b):

All distinct permeable zones (group of zones originally at the same pressure regime

between which fluids can flow) through which the well passes must be isolated from

the surface, and from each other, by at least one permanent barrier (i.e., a cement

plug). The barrier must provide a permanent seal and extend across the entire cross-

section of the well.

An additional barrier should be installed at the base of any groundwater formations.

This is designed to prevent contamination of groundwater in addition to protecting

surface waters and soils from any fluids within the well. Many US States require

barriers to be installed at the base of the surface casing and / or between each

production zone.

If the permeable zone is hydrocarbon bearing then two barriers are required, although

these can be combined into a single large barrier covering the entire vertical extent of

the hydrocarbon bearing zone.

In the case where multiple hydrocarbon bearing zones are in close vertical proximity

to each other, the lower barrier for the upper zone can act as the upper barrier for the

lower zone.

74

Cement is the material most commonly used to form the barriers. Other materials

can be used although they must have certain properties, i.e., permeability, integrity,

non-shrinkage, ductility versus brittleness, chemical resistance and bonding capacity.

Permanent barriers are recommended to be at least 30 m in length; however, a barrier

length of 150 m should be used where possible.

If permeable zones are <30 m apart, it is recommended that the gap between the

zones be completely bridged with a cement column. The top of the upper section of

the barrier should extend >30 m above the highest potential flow point.

In the situation where two barriers are combined into one, the total length should be

>60 m. More typically, a barrier length of 250 m is used. The top of the barrier should

extend 60 m above the highest potential flow point.

The integrity of the annulus directly adjacent to the cement plug must be ensured in

order to form a complete seal. The original cement and casing of the well are not

considered to be permanent barriers as they can corrode or degrade with time. If

problems with the cement or casing are detected prior to the installation of the

barriers, remedial action must be taken. Such action can include perforating the

casing and pumping cement slurry through the perforations (cement squeeze) or

milling away casing and / or cement and replacing the removed material with cement.

Permanent barriers, once installed, must be tested and verified in order to ensure that

they are in the correct location and form an adequate seal with the well casing.

If a zone of irretrievable radioactive material is found, a permanent barrier should be

installed and the zone sealed off. These areas should also be surveyed, tagged and

reported to the appropriate authority.

Removal of downhole equipment is not required as long as the seal formed by the

installed barrier is sufficient and zonal isolation is achieved. However, wires and / or

cables should not form part of the barrier as they can act of potential flow paths.

If well completion casing, e.g., production casing, cannot be removed from the well,

extra precautions must be made to ensure that the barriers around such areas are

sufficient and that the performance of the barriers is verified.

Barriers installed in wells with significant concentrations of acidic compounds, e.g.,

CO2 and H2S, the barriers should be designed to withstand corrosion from such

compounds.

Potentially harmful or polluting fluids, e.g., mud and hydrocarbons, found above the

uppermost barrier should be removed as much as is reasonably possible. A cement

plug should then be installed at / near to the ground surface to prevent runoff entering

the well.

Finally, the casing should be cut-off below the ground surface. A casing stub (plate)

should then be welded onto the top of the well. The well cellar should then be filled

with a suitable material and surface condition restored to a level close to that prior to

drilling.

75

The EA also outline problems that can occur during, and after, well abandonment. The two

main challenges are the inability to remove down-hole equipment, the method of dealing

with this is outlined above, and the failure to create adequate seals (Environment Agency,

2012b). The main problem that could potentially be encountered after abandonment is

corrosion resulting in the connection between flowing zones (Environment Agency, 2012b).

Once a well has been abandoned to a satisfactory degree, the responsibility of aftercare and

monitoring lies with the operator (United Kingdom Onshore Operators Group, 2015).

From the research carried out for this report, there appears to be no provisions or legislation

in place regarding what happens to abandoned decommissioned and abandoned wells if the

operator goes out of business. This is a severe shortcoming in the current legislation that will

need to be addressed if shale gas exploration and extraction is to go ahead in the UK.

Currently, there is little in the way of monitoring of abandoned wells in the UK (Davies et al.,

2014). These same authors proposed that the wells should be inspected 2-3 months after the

cement plugs have been inserted into the well. Inspection should cover gas migration and

casing pressure. Inspection frequency can then be reduced providing that there is no

evidence of the well leaking; the well can then be cut off and buried. Any further inspection

should focus on soil monitoring, at a recommended interval of once every five years (Davies

et al., 2014) or if there is cause to think that well integrity might be compromised, e.g., a

seismic event. DECC, together with industry operators, are currently putting in place rules to

ensure that monitoring and restoration are maintained at abandoned sites, even if the

operator goes out of business (Department of Energy and Climate Change, 2014b).

76

Waste residues

During the drilling of the well, cuttings are produced and must be disposed of in an

environmentally friendly manner. The most common methods of disposing of the drill

cuttings is either transport to hazardous or non-hazardous landfill site, depending on the

concentration of salt and other chemicals, or spreading on agricultural land, both of which

are normally permitted by the EA (Stamford and Azapagic, 2014). However, spreading onto

agricultural land is not normally carried out in the UK. Some of the waste may be

contaminated, either by material dissolved from the shale, from the fluid base (i.e., oil) or

from chemicals in the fluid. In this case the method of disposal will need more careful

consideration. Stamford and Azapagic (2014) suggested that in the average well, 60% will be

spread on agricultural land while the remaining 40% will be disposed of in landfill. In the best

case scenario, 100% would be deposited in landfill, while in the worst case, all the waste will

be spread over agricultural land.

The main form of waste resulting from hydraulic fracturing is the water that returns to the

surface after fracturing has taken place, i.e., flowback fluid. 20-40% of the fluid injected into

the well returns to the surface as flowback following the fracturing stage, the majority of the

remaining fluid returns to the surface over time (Cuadrilla Resources Ltd., 2015). Note that

the fluid returns through the well, not the surrounding rock. Approximately 80% of the fluid

introduced to the well returns to the surface (note this is inclusive of the 20-40% that returns

during the fracturing stage), the remainder remains in the fractured rock. Due to the lack of

data on hydraulic fracturing of shale in the UK and the variability of local geology, it is difficult

to say how much of the fluid will return to the ground surface (Stamford and Azapagic, 2014).

As discussed previously, once these fluids have returned to the surface, there is a risk of

surface contamination if a spillage occurs. This is particularly the case in the US where open

pit storage is commonly used to store flowback fluid. In order to prevent any surface

contamination, UK regulation dictates that the fluid must be safely stored in robust covered

steel tanks before being treated and disposed of (Department of Energy and Climate Change,

2014e). This is done in order to minimise any emissions to the atmosphere and reduce the

risk of spillage. The drilling site must also be designed so that any spillage is avoided. It should

be noted that the use of tanks does not eliminate emissions as, as the tanks fill up, air and

hydrocarbon VOCs are vented to the atmosphere. In order to minimise emissions, Green

Completions are used. This allows gas that would be vented is either safely flared or captured

and sold (see Mitigating and controlling emissions section for more details regarding green

completions). In the US, the use of green completions is now mandatory during the

production stage but not for the exploration and appraisal stages. If a spill does take place,

measures must be taken to contain the spill and minimise the environmental impact.

The flowback fluid itself is classified by DECC as mining waste and, as such, needs to be

disposed of in accordance to the mining waste regulations put in place by the relevant

national EA. In addition, a plan for waste management based on laboratory tests must be in

77

place before drilling begins. During the operation of the site, the relevant authorities must

be informed before any movement of waste takes place (Department of Energy and Climate

Change, 2014e).

There are a number of methods by which flowback can be suitably disposed. One such

method is on-site recycling where the flowback is treated and the water re-used in the

fracturing process (Rassenfoss, 2011). The remaining material that cannot be recycled is

transported to a waste treatment facility. Another method is simply to transport all of the

waste from the site to an external treatment and disposal facility. Lastly, the waste material

can be fed into a special sewer designed to handle waste water. This method requires special

permission from the local utilities company (Department of Energy and Climate Change,

2014e).

The radioactive elements brought back to the surface in flowback fluid pose a potential hazard

to the site employees if the concentration of NORM is high or if prolonged exposure occurs

(Environment Agency, 2013). The way of dealing with the NORM brought up to the surface is

dependent upon the concentration or activity in the flowback. If the concentration is above

thresholds previously defined by the EA, a radioactive substances licence is required, whereas

if the concentration is below the threshold, the waste will be disposed according to those

defined by mining waste regulations. As with flowback disposal, the operator must submit a

plan for the disposal of the NORM contaminated material that includes details about the

handling and disposal process, in addition to the measures in place to mitigate the risk to

people and the environment (Department of Energy and Climate Change, 2014e).

At the Preese Hall site, Cuadrilla stored all of their wastewater in double-skinned steel tanks

(Environment Agency, 2011; Cuadrilla Resources Ltd, 2015). Both Cuadrilla and the EA carried

out regular tests on the wastewater; note that the EA can also carryout unannounced tests.

Once the water was ready for disposal, the wastewater was transported to a water treatment

plant in Davyhulme (Environment Agency, 2011; Cuadrilla Resources Ltd, 2015). This

treatment plant currently treats wastewater from other industrial sources and has the

capability to deal with the contaminants within the wastewater (Environment Agency, 2011).

Before treatment, the water is tested by the treatment plant (Cuadrilla Resources Ltd, 2015).

Once treated, the Davyhulme plant is permitted to discharge the water into the Manchester

Ship Canal (Environment Agency, 2011). A general overview of the water treatment process

in the UK can be seen in Figure 24.

78

Figure 24. Diagram displaying the stages of water treatment associated with shale gas operations in the UK. (From personal communication with Mike Holgate).

There is currently much research taking place into alternative fracturing fluids that do not

dissolve NORM (The Royal Society and the Royal Academy of Engineering, 2012), metals and

minerals, therefore reducing the hazard associated with these substances. However, these

fluids are not yet commercially available (Stamford and Azapagic, 2014).

Research carried out as part of the ReFINE independent research consortium has determined

that the level of NORM, specifically potassium (K-40) and radium (Ra-226 and Ra-228), in

flowback fluid, although higher than that present in groundwater, are below the permitted

UK exposure limits (Almond et al., 2014). The maximum recommended annual personal dose

above natural background levels in the UK is 1 mSv (millisieverts) (Almond et al., 2014). Text

box 2 gives a basic overview of the types of radioactivity.

79

The Almond et al. (2014) study analysed the flowback fluid expelled from the Carboniferous

Barnett Shale (US), the Carboniferous Bowland Shale (UK), and the Silurian Shales of Poland.

To provide a reference for the concentration of NORM in flowback fluid, groundwater from

the basins containing each of the three shales was also analysed, as were surface water

sources, i.e., reservoirs. Background levels of both radium species measured in groundwater

Text box 2 – Types of radioactivity

There are three different types of radiation, i.e., alpha, beta and gamma. Each type of radiation can travel

different differences and can penetrate different materials.

Alpha particles are simply protons, and examples of alpha particle emitters are thorium-232, radon-222,

uranium-238 and radium-226. The latter two are present in varying amount in nearly all rocks, soils and

water (US Environmental Protection Agency, 2012). The particles travel a few centimetres from their source

and can be stopped by a piece of paper or clothing, therefore, in order for human exposure to occur, alpha

particles must be ingested (Classic, 2013). If inhaled, alpha particles have been shown to cause lung cancer

(US Environmental Protection Agency, 2012).

Beta particles are simply electrons and can travel short distances up to 1 m and is moderately penetrating,

however, particles can be stopped by a 3 mm piece of aluminium but can penetrate clothing and skin down

to the level where new skin cells are produced (the germinal layer) (Classic, 2013) If prolonged skin exposure

to high levels of beta-emitting materials takes place there is the potential to cause skin injury. Examples of

beta emitters are strontium-90, carbon-14, sulphur-35 and radium-228. The latter of which can be harmful

if ingested (Sperger et al., 2012; Classic, 2013).

Gamma radiation is a form of electromagnetic radiation similar to X-rays but are of comparatively higher

energy, higher frequency and shorter wavelength (Classic, 2013). They are also the most energetic type of

radiation depending on the intensity of the source, can require several inches of dense material, e.g., lead

or metres of cement to absorb the waves. This means that gamma rays can travel through the body,

therefore making them the most dangerous of the three types of radiation (Classic, 2013). Some examples

of gamma emitters are iodine-131 and radium-226. .

Radiation is measured in either grays (Gy) or sieverts (Sv) for absorbed and equivalent dose respectively.

Note that these units differ from Becquerels (Bq) which are a measure of radioactivity given in terms of the

number of decays per second.

Many radioactive elements occur naturally as part of rocks the in the Earth’s crust, e.g., uranium, radon and

potassium, as such, they are included in the NORM umbrella term (Sperger et al., 2012). Rocks, including

shale, contain some of these radioactive elements. For instance, the Marcellus Shale contains naturally

occurring uranium and thorium and their radioactive decay products, e.g., radium-226 (Sperger et al., 2012).

Because rocks contain radioactive elements, the risk these pose will depend on the elements in the rock and

their concentrations, and hence the type and amount of radiation emitted.

The concentration of uranium within shale may mean that it emits 20 times the background level of

radiation, although this still a tiny amount. Because of this, gas-bearing shale deposits have been located

based on detected gamma radiation levels (Resnikoff et al., 2010). The detection of gamma radiation is also

used to detect the total organic carbon (TOC) content of shales beneath the ground surface.

80

were found to be higher in the Silurian and Barnett shales compared to the Bowland shale

(Table 8). For the flowback fluids, 1% exceedance levels (the flux expected to be exceeded

1% of the time, i.e., a worst case scenario) were determined. For US and UK shales, this was

found to be approximately 0.09 mSv and for Polish shales the value was 0.43 mSv (ReFINE,

2014). Levels of radiation in both US and Polish groundwaters were 7-8 times lower than in

flowback waters (ReFINE, 2014). In the UK samples, the flowback fluid concentration was

approximately 500 times higher than the groundwater. However, the concentration in

flowback fluid was much lower than in flowback fluids from the conventional oil and gas

industry where there is a 50% likelihood that that radioactivity will exceed 13 mSv (compare

this to the 0.09 mSv of UK and US shales) (ReFINE, 2014). NORM output from nuclear power

station discharge waters was found to be two orders of magnitude higher than that of

flowback fluids produced during shale gas operations (Almond et al., 2014).

Table 8. Table displaying the upper and lower concentrations of radioactive species measured in groundwater of the Polish Silurian shales, the Barnett shales, the Bowland shales and an amalgamation of measurements taken from across the World. n is the number of samples. Also of note is the small number of samples analysed from the Bowland Shale on account of the lack of drilling activity in the UK. (From Almond et al., 2014)

With regard to the Bowland Shale specifically, K-40 (a beta emitter, therefore considered to

be analogous to levels of beta radiation) in flowback fluid was higher (112,958 Bq/yr) than

that of groundwater (4,870 Bq/yr). The same was also the case for Ra-226 (an alpha emitter,

therefore considered analogous to alpha radiation) where flowback contained

2,379,175 Bq/yr while groundwater contained 1,460 Bq/yr. Note that although higher than

groundwater, the radioactivity of the flowback presents only a localised hazard due to this

being alpha radiation.

When compared to global groundwater levels of K-40 and Ra-226, the increased radioactivity

of flowback fluid is put into perspective. For instance, the global distribution of Ra-226 in

groundwater was determined to be 1,650,000 Bq/yr which is three orders of magnitude

greater than the radioactivity of groundwater overlying the Bowland Shale but lower than the

levels in flowback fluid (Almond et al., 2014). Global K-40 radioactivity levels were calculated

as being five times higher than those of UK groundwater (Almond et al., 2014). Additionally,

81

the total surface water radioactivity was two orders of magnitude lower than that of the

flowback fluid (approximately 21,000 Bq/yr compared to 2,379,175 Bq/yr) (Almond et al.,

2014). The authors use this difference to highlight the fact that contamination of surface

water sources by flowback fluid could pose a significant risk.

The radioactivity in flowback was compared to other anthropogenic sources, specifically

discharge water from nuclear power stations and waste water produced by the conventional

hydrocarbon industry. In order to provide a more meaningful comparison between shale gas

flowback fluid radioactivity and radioactivity from other industrial sources we can compare

the “radioactive footprint of the energy source”; this is determined as Bq/kW/h. The largest

radioactivity footprint is associated with nuclear power generation (28,740 Bq/kW/h)

followed by coal (374 Bq/kW/h). For the latter, the main radioactive source is burn products,

e.g., ash. Shale gas would be the third largest (7 Bq/kW/h (based on 1% exceedance levels)).

The Almond et al. (2014) concluded that, although the levels of NORM in flowback fluid are

higher than those found in groundwater, they are below the safe limit levels of exposure

(ReFINE, 2014); in no scenario did the 1% exceedance exposure, i.e., a worst case scenario,

exceed the allowable annual exposure. In comparison with other industries, the level of

radiation in flowback fluid is lower (ReFINE, 2014). Based on these two findings, the authors

conclude that the fluids are unlikely to pose a threat to human health (ReFINE, 2014).

As an aside, Almond et al. (2014) provided figures that allow the waste water produced during

shale gas operations to be put into some form of context. They say that, in 2007,

400,000,000 m3 (400 million tonnes) of water produced by oil operations was treated and

discharged into the North Sea; an additional 85 m3 was injected into the subsurface. In 2010,

there were 104 UK installations discharging an annual volume of 196,333,229 m3 of water

into the sea, 27,481,713 m3 of which was produced water (Almond et al., 2014 citing OSPAR

Commission, 2012). Based on the estimated volume of flowback fluid produced from the

Preese Hall drilling (6,627 m3), the authors hypothesise that if development took place in the

UK at the rate of 25 wells a year for 20 years, the maximum expected volume of flowback

fluid would be 3,313,500 m3. Even with 250 wells per year for 20 years, the volume of

flowback would be 33,135,000 m3 which is still much less than that produced by oil

operations. Almond et al. (2014) consider this as an indication that the UK already has

experience with treating and disposing of much larger volume of waste water. It should,

however, be noted that the study does not address the issue of how this water would be

obtained or if extracting such volumes of water would put strain on the UK domestic water

supply.

82

Air

A potential risk to human health may arise from emissions originating from the drilling site.

Emissions can emanate from two main sources, either:

a) fugitive emission of methane and other volatile organic compounds or

b) from machinery working on-site and vehicles transporting materials to and from the site.

Fugitive emissions can be produced by loss of well integrity, resulting in leaks. Emissions can

also be produced from flowback fluid, if not properly stored, and from the process of

compression, condensation and transportation of methane (McKenzie et al., 2012).

MacKay and Stone (2013) conducted a report for DECC on the greenhouse gas emissions

associated with shale gas. They pointed out that there is great uncertainty regarding actual

emissions from shale gas operations due to the currently limited available data concerning

direct measurements of emissions. What figures that are available are based on engineering

calculations and approximate measurements of gas flow (MacKay and Stone, 2013). A recent

study has demonstrated that estimates of emissions in the US, made in 2011 by the

Environmental Protection Agency (EPA), were higher than the actual emissions output from

natural gas sites (Allen et al., 2013), although the figures are in reasonable agreement with

each other. Overall estimates were put at 2,545 Gg methane/yr (note; 1 Gg = 109 g) while

actual measurements were put at 2,300 Gg methane/yr. However, some of the individual

sources of emissions were larger than estimated (e.g., leaks from pneumatic controllers) and

some were smaller than estimated (e.g., emissions associated with flowback completion). It

should be noted that some of the emission sources (e.g., work overs) were not sufficiently

well measured as to allow accurate data to be acquired. This, in conjunction with the

resolution of the measuring equipment, means that caution must be used when applying the

results.

MacKay and Stone (2013) also pointed out that the environmental impact statement, which

is required to be submitted by operators to the EA and Mineral Planning authority, must

contain details about the site emissions expected by the operator. The EA will then review

the expected emissions and, if necessary, can enforce monitoring procedures. To ensure that

the risk to the environment is minimised, making monitoring mandatory, no matter what the

expected emissions, should be considered. The authors also recognise that because the shale

gas industry in the UK is in its infancy, comprehensive path finder studies, aimed at

establishing which emissions come from where and how much is produced, should be put in

place at a small number of sites. The resulting information can then be used as a basis for

monitoring at future sites.

83

Monitoring air quality and emissions

The EA (Broomfield and Donovan, 2012) published a comprehensive document on the

monitoring and control of fugitive emissions coming from shale gas operations. Note that the

vast majority of the information in the following sections on detection and monitoring

techniques is taken from this study.

Broomfield and Donovan (2012) also highlighted that the monitoring requirements at

different sites will change with time. Prior to drilling it is essential that background levels of

methane are established. Once drilling begins, the focus of the monitoring will likely shift

towards the monitoring of fugitive emissions through on-site, fence line and regional

monitoring aimed at assessing the methane flux at the site.

Broomfield and Donovan (2012) pointed out that if the amount of methane being emitted

from a leak, in terms of financial worth, is less than the cost of the repair, then the repair is

generally not carried out. This approach is not acceptable if, as the Environment Agency

expects, 100% of emissions that come from shale gas sites will be contained (Environmental

Audit Committee, 2015, paragraph 52). Thus all leaks will have to be addressed, not just from

the point of view of the financial loss. Broomfield and Donovan (2012) do not address the

industry approach to multiple small leaks that are not releasing enough methane to make

repair financially viable. It is possible that large fugitive emission levels could result from

many small leaks that the operator deems individually to be not financially viable to fix.

However, it must be anticipated that in future if this did occur the operator would be

compelled, either by social responsibility or by legislation, to carry out remedial action.

As an aside, the question must be asked, how attainable is the EA’s target of zero fugitive

emissions? For instance, is it financially viable for operators to mitigate completely all

emissions? One would assume that in order to meet such a target, real-time monitoring or

some similar form of intense monitoring will need to be permanently installed on-site in order

to identify any leaks immediately. Another option would be for a team constantly to be

available on-site to detect and deal with leaks. Both of these might prove to be prohibitively

expensive. Also, it is unclear as to whether this target is inclusive or exclusive of emissions

that occur during the everyday operation of components. For instance compressors release

emissions as part of their design and, although the amount of such emissions can be

mitigated, it cannot be eliminated completely. Therefore, if the EA consider zero emissions

literally to mean zero emissions, the goal may be unattainable and should be recognized as

such.

Leak and emission detection techniques

The first step carried out prior to detecting leaks is establishing the areas of the site that are

most likely to produce fugitive emissions. These are commonly considered to be storage

tanks, seals, valves, compressors etc. The entire site is assessed and a complete inventory of

84

potential leak sites and the associated hazard calculated. This process is known as “leak

detection and repair” (Broomfield and Donovan, 2012). It can be used as a basis for the

placement of monitoring equipment.

Once a leak detection protocol has been established, the process of leak detection begins.

When a leak is found, it can take anywhere between 48 hours and 15 weeks to repair it

(Broomfield and Donovan, 2012). If the leak is significant, i.e., such that repair would require

disruption of the operation, then the repair can be put on hold until such a time where

minimal disruption will take place, i.e., during a shutdown (Broomfield and Donovan, 2012).

This would not be compatible with the enforcement of emissions policies, driven not by safety

considerations, but by the desire to minimise greenhouse gas emissions.

In order to detect leaks, and the resulting emissions, a number of different pieces of

equipment are available to operators. These include: hot bead and catalytic combustion

analysers, forward looking infrared, infrared adsorption spectroscopy, flame ionisation

detection and non-dispersive infrared detection.

Hot bead and catalytic combustion analysers are the most commonly used lower explosive

level (LEL) detection techniques. LEL is defined as the minimum amount of gas needed to

cause an explosion when ignited in the presence of oxygen. In the case of methane, the LEL

is 5 vol%. However, at regular surface conditions (20 °C and 1 atmosphere pressure) the

hazard is greatest at 9.5 vol% methane. This allows the explosion risk of any gas leak to be

determined (RKI Instruments, 2014). It should be noted that the LEL measurement does not

apply solely to methane, but also to other hydrocarbons emitted from a site that can mix with

methane to form an explosive mixture (Broomfield and Donovan, 2012). The HSE has

produced a document detailing such factors as the range, selection criteria, inspection,

maintenance and calibration of the LEL detectors (Health and Safety Executive, 2004).

LEL detection equipment is relatively low-cost (£1,000-3,000 each), portable, safe to use, and

durable. Methane concentration is measured by comparing the resistance in a Wheatstone

Bridge circuit, in which one of the arms of the circuit has a catalytic substrate, the other a

reference substrate. The presence of combustible gas, i.e., methane, will ignite the catalytic

substrate which changes the resistance characteristics of the circuit. The change is

proportional to the concentration of the combustible gas (Broomfield and Donovan, 2012).

Forward looking infrared (FLIR) and infrared absorption spectroscopy (IAS) are becoming

more popular methods of detecting leaks. The FLIR technique allows real-time imaging of

fugitive emissions by way of a small screen mounted on a hand-held device, allowing the

operator quickly to identify the origin and magnitude of leaks. The technique is, however,

very sensitive to weather conditions. IAS uses a semiconductor laser (tuned diode laser –

TDL) which is fired through the area of interest. The device then measures the absorption in

a particular wavelength range in the beam reflected from a target. The amount of absorption

correlates with the methane concentration in the beam path. This technique is not sensitive

85

to weather conditions, and therefore has an advantage over FLIR. However, IAS does not

provide an image of the emission magnitude and source. Unlike FLIR it must also be pointed

directly at the emission source. These two techniques are therefore presently best used in

conjunction with each other but new technical developments are expected shortly to expand

significantly the utility of these devices. Prices vary between £1,500-50,000 for these devices

(Broomfield and Donovan, 2012).

Flame ionisation detection (FID) has been the most widely used method of fugitive emissions

monitoring. In this technique, the sample chamber in the device contains a flame fuelled by

hydrocarbon-free air and hydrogen. When the external air sample is introduced in to the

chamber, any methane present will be ionised into carbon, which changes the electric current

flowing across the chamber by an amount proportional to the amount of methane in the

sample. VOCs can also be measured using this same technique. The main drawback with this

technique is the fact that an open flame is used, therefore making the method less safe

compared to LEL, IAS and FLIR techniques. The devices used to conduct the measurements

are also sensitive to other types of hydrocarbons. In terms of cost, hand-held devices cost

between £1,600 and £6,000 with higher-end devices that are capable of detecting only

methane can cost between £9,000 and £16,000 (Broomfield and Donovan, 2012).

Non-dispersive infrared detection spectroscopy (NDIR) uses an infrared beam, the

wavelength of which is attuned to that absorbed by methane. When the infrared signal

detected at the end of the sample chamber is compared to that of an infrared beam in an

inert gas, a measurement of the methane content of the gas can be made. The equipment

costs in the region of £6,000 to £10,000 (Broomfield and Donovan, 2012).

Discrete ambient air measurements

Discrete ambient measurements can be made in order to determine methane, as well as VOC

concentrations. The technique for obtaining these measurements is first to collect air

samples in stainless steel canisters, known as Summa canisters. These are analysed using a

two-step process. First the components of the gas are separated using gas chromatography,

the products of which are then quantified using mass spectrometry. This technique allows a

snap-shot of the air composition the time of sampling (Broomfield and Donovan, 2012). It is

particularly useful as a first approach in emergency situations where the air composition

needs to be quickly established (Broomfield and Donovan, 2012). This technique is relatively

cost-efficient as canisters are low-priced and lab costs start at approximately £70 per sample

(Broomfield and Donovan, 2012). However, if a large number of samples are to be analysed,

i.e., from across an entire site, this cost will increase.

Cavity enhanced adsorption spectroscopy is another method of assessing ambient air

composition that produces higher accuracy results compared to the sampling method

described above (Broomfield and Donovan, 2012). This technique involves a laser being fired

86

into an absorption cell (i.e., the cavity) containing at least two mirrors. The laser beam reflects

off these mirrors multiple times, usually a total length of >10 km. As the mirrors are not

perfect, the intensity, and therefore energy, of the laser decreases with time, eventually

reaching zero. The laser itself can be attuned to be absorbed by the gas of interest; once the

gas is introduced into the chamber, the amount of time required for the energy of the laser

to reach zero decreases. The difference between the amount of time required for the laser

to stop reflecting in the sample-free chamber and the chamber containing the air sample is

proportional to the concentration of the gas of interest (Broomfield and Donovan, 2012). The

main drawback of this technique is the cost. Although the technique is sensitive to 1 ppb

(parts per billion) or less, the cost for portable devices is around £35,000, with lab-based

devices being around £27,000 (Broomfield and Donovan, 2012).

Open source and whole site fence line monitoring

Fence line monitoring systems are used around the boundaries of drill sites. Monitoring

systems can be placed further afield in order to measure the wider distribution of emissions.

Fence line monitoring can be used to determine the methane flux across a site. This is done

by setting up two sets of monitoring stations, one on the upwind side of the site and one on

the downwind side of the site. The upwind stations allow a background emission

measurement to be made. Additionally, if monitoring stations are placed around a region

containing multiple wells, the regional methane flux can be calculated (Broomfield and

Donovan, 2012). As this is, effectively, a large scale monitoring process, it is necessary that

fence line monitoring be used in combination with other monitoring methods that determine

the fugitive emission levels directly from potential sources of fugitive emissions, such as

flowback storage tanks.

Four main techniques are available to detect the concentration (i.e. areas of large fugitive

emission concentration, known as hotspots) and flux of emissions across drilling sites. These

are (Broomfield and Donovan, 2012):

Open path Fourier transform infrared (OP-FTIR)

Ultraviolet differential optical absorption spectroscopy (UV-DOAS)

Tuneable diode laser absorption spectroscopy (TDLAS)

Path integrated differential absorption spectroscopy (PI-DIAL)

These systems are expensive to implement, however, they are capable of mapping the

concentration across large areas and, therefore, there is the possibility of the large one-off

cost of these techniques being off-set by the cost of multiple cheaper, but less accurate,

devices. Additionally, the measurements made using these techniques can be combined with

87

statistical and computer models in order to determine the location and size of emissions

sources at ground level, e.g. emissions from storage tanks.

In basic terms, these techniques measure gases along linear sections between a transmitter

and receiver. Radial plume mapping is a method applied using the above techniques that

allows the identification of emission hotspots (Broomfield and Donovan, 2012). In order to

identify hotspots, a series of measurements is made across the area of interest in the

horizontal plane. This creates a concentration contour map of the gas of interest. Another

series of measurements are made vertically, this creates a cross-section of the emissions

plume. The two together effectively create a 3D-image. If the distance between the receiver

and transmitter is reduced, more detailed results can be obtained. The equipment is

commonly set up downwind of the site to allow hotspots to be identified (Broomfield and

Donovan, 2012).

As a side note, because measurements are made over the entire site, details about the

meteorological conditions at the time at which the measurements were taken are also

required (Broomfield and Donovan, 2012). If these techniques are applied, the operator

should install a weather monitoring system on-site. The station must be capable of

establishing the variation in wind speed and direction in both the horizontal and vertical

planes (Broomfield and Donovan, 2012). The air pressure and turbulence, in addition to the

amount of solar radiation, humidity and the dew point, must also be quantified (Broomfield

and Donovan, 2012).

The above discussion relates to the most common methods of radial plume mapping. A

variation on these methods comes in the form of LIDAR (Laser Illuminated Detection and

Ranging) based plume mapping using path integrated differential adsorption (DIAL). In this

method, a dual-wavelength beam is projected over the area of interest. One wavelength is

attuned to be absorbed by the methane and the other one not. The difference between the

returning signals is proportional to the methane content in the examined area. This technique

provides high resolution, real-time measurements attainable over a short timescale.

However, the main drawbacks are cost and limited availability. For instance, there are only

two systems in the UK, both of which are owned by the National Physical Laboratory. These

are truck-mounted mobile systems, with a detection range of up to 3 km and a spatial

resolution of <8 m, that could be brought on-site periodically to make detailed

measurements. The operators could then rely on lower cost, lower detail techniques for the

remainder of the time. New DIAL systems can cost more than £500,000 each, therefore it is

not likely that individual operators in the UK will buy these systems. A potentially feasible

plan would be for the UK operators to share the cost of purchasing one or more of these

systems. The National Physical Laboratory charge approximately £30,000 per site visit for the

DIAL systems. Again this is a considerable cost for an operator for a one week long survey.

However, the approach of installing lower accuracy detectors and periodically hiring a DIAL

88

system could be more economically viable compared to the cost of installing or buying a

number of smaller, high accuracy devices.

The technique of tracer gas correlation can be used with the previously discussed methods in

order to obtain more accurate measurements of methane flux. This method requires the

controlled release of a known gas at a known flow rate. The concentration is measured con-

currently with methane measurements. In addition, knowledge of the emission source is

required for this technique.

Mitigating and controlling emissions

The majority of fugitive emissions are released after the hydraulic fracturing stage as the

flowback returns to the surface (Department of Energy and Climate Change, 2014a). A

proportion of the flowback is made up of hydrocarbons and natural methane gas. In order to

contain and process the flowback, equipment known as Reduced Emissions Completions, or

green completions, is used. This treatment process acts to separate the gas and hydrocarbons

from the remaining flowback to allow the gas and hydrocarbons to be contained (up to 90%

of the gas is recovered) and the remaining fluid to go on to further processing (IPIECA, 2013).

This process also reduces fugitive emissions by 90% (from 312,000 m3 to 31,000 m3) (Stamford

and Azapagic, 2014). Recent policy changes in the US have mandated the use of green

completions; therefore, it is likely that the same policy will be adopted in the UK (Stamford

and Azapagic, 2014).

In addition to green completions, operators can also manage methane loss through two other

methods. Firstly, the gas can be vented which involves the controlled release of gas into the

atmosphere without burning. By not burning the methane, the level of greenhouse gases

output by the site increases. As a result, venting may only be carried out when there is a

safety hazard if the gas were to be ignited (Department of Energy and Climate Change, 2014a).

The second method involves controlled on-site burning, or “flaring”, of the methane. This

reduces greenhouse gas emissions by approximately 80% when compared to venting

(Department of Energy and Climate Change, 2014a). Flaring has been documented to have

an adverse effect on human health. It has been shown that people living in communities

where gas flaring takes place in Nigeria have reduced lung function. The more prolonged the

exposure, the larger the reduction in lung function (Ovuakporaye et al., 2012). However, it

should be noted that Nigeria flares and vents more gas than any other country (19.79% of the

total global flaring in 2001) (Friends of the Earth, 2005). Friends of the Earth (2005) estimated

that the between 2-2.5 bcf (billion cubic feet) of gas was flared per day, this equates to

approximately 25% of the UK’s annual gas consumption. Therefore, the level of flaring that

may take place in the UK would be minute in comparison.

It should be noted that it is not in the interest of the operator to vent or flare the gas, not only

for environmental reasons, but also for economic reasons. Any gas that is lost translates to

89

potential lost profit. Therefore, one should expect the operator to take as many steps as

possible to minimise the methane loss. There is currently no data regarding the amount of

gas vented in the UK (Stamford and Azapagic, 2014).

In the US, the use of green completions has reduced the need for flaring by 70-90% (IPIECA,

2013). There are other alternatives that can be used to reduce the reliance on flaring. It

should be noted that during the research carried out for this briefing paper, no evidence of

flaring being completely eliminated was found. One technique designed to reduce the

reliance on flaring involves using the initially produced gas to generate power, either on-site

or in the local communities (Gibson, 2013). Another method that has been used in the

conventional oil and gas industry is re-injection of the gas into the wellbore annulus to

facilitate gas lift (Gibson, 2013). Gas lift is the process by which re-injected gas lifts the well

fluids to the surface, this allows the production rate from the well to be increased (Petrowiki,

2015). If a pipeline is not present at the site at the time of exploration, when flaring is most

likely to be used, compressing the gas to produce Liquefied Natural Gas (LNG) or Gas to

Liquids (GTLS), e.g., methanol or dimethyl ether, can be considered as an alternative to flaring

(Gibson, 2013). Flare Gas Recovery (FGR) systems are used when the site is operating under

low or normal pressure conditions. This allows the flares to be shut down and any gas that

would normally be flared can be collected, compressed and re-routed for other off-site uses

(Gibson, 2013). These have been installed at conventional hydrocarbon sites and have

reduced the need for flaring to “near zero” (Gibson, 2013). The two main drawbacks for this

technique are the cost and space requirements (Gibson, 2013); therefore the viability will

need to be assessed on a site by site basis.

An additional method that is considered to be the equivalent burning gas through flaring is

through the use of an incinerator. In the conventional hydrocarbon industry, these are most

commonly used at sour gas processing plants where natural gas and / or hydrogen sulphide

(H2S) must be disposed of (Bott, 2007). It should be noted that these differ from enclosed

flares which are simply flares protected from the weather (Bott, 2007). Enclosed flares are

not “true” incinerators as the latter require controls to maintain specific air-to-fuel ratios, a

refractory lining and a minimum residence time (the time the gas spends in the combustion

chamber before being released into the atmosphere) for the gas (Bott, 2007). Despite this,

incinerators can generally be thought of as flares contained within a combustion chamber,

indeed it is common practice not to differentiate between the emissions from flaring and

incineration, they are usually considered under the umbrella term of “flaring” (Bott, 2007).

Using enclosed flares and incinerators has multiple advantages over conventional, exposed

flares. Flares are no longer visible, thereby reducing light pollution (BC Oil and Gas

Commission, 2011). Additionally, using an incinerator allows more efficient control over the

combustion process as the amount of oxygen that is added to the chamber can be controlled

(BC Oil and Gas Commission, 2011). The use of an incinerator does not necessarily result in

reduced CO2 emissions. The amount of CO2 produced by both flaring and incineration is

considered to be approximately equal (BC Oil and Gas Commission, 2011).

90

In terms of a case study, British Columbia carried out baseline measurements of flaring of gas

at hydrocarbon wells in 2006. Between 2006 and 2009, the volume of gas being flared

annually dropped by 23%. The volume of gas being flared at oil production wells and oil

batteries has decreased by 92% since 1997 (BC Oil and Gas Commission, 2011). As of 2011,

96% of extracted gas is captured, either for sale or used for other purposes, e.g. on-site power

generation. In total, flaring of gas accounts for <2% of the greenhouse gas emissions in British

Columbia (BC Oil and Gas Commission, 2011).

It is worth considering, in a more general sense, the type and amount of emissions produced

during flaring and incineration. The objective of both of these processes is to convert as much

methane as possible into CO2 and water vapour. The degree to which this occurs is known as

the combustion efficiency (Alberta Energy Regulator, 2014). A reduction in combustion

efficiency can result in hydrocarbons being released into the atmosphere. Combustion

efficiency can be influenced by a number of factors, including meteorological conditions,

operator competency and waste gas composition (Alberta Energy Regulator, 2014). Windy

conditions will interfere with an exposed flare, resulting in a reduction in efficiency, whereas

incinerators have no such problem. However, because open flare stacks are taller than

incinerators, the products of the flaring tend to be better dispersed into the atmosphere

(Alberta Energy Regulator, 2014). This is a minor problem when combustion efficiency is high

but low efficiency results in methane being released in to the atmosphere and not being well

dispersed, though the greenhouse gas impact remains the same. If the gas contains other

substances (e.g., H2S which is converted to sulphur dioxide (SO2) when burnt) the hazard

posed by poor dispersion of incineration products is larger (Alberta Energy Regulator, 2014).

In order for an incinerator to work efficiently, the equipment must operate at a sufficiently

high temperature and the gas must have a sufficient residence time. For instance, in Alberta,

Canada, the Alberta Energy Regulator (2014) require incinerators to operate at a minimum

temperature of 600 °C and have a residence time of at least 0.5 seconds. The same

Department requires flares to have a minimum gas energy content of 20 MJ/m3, if the gas

does not meet this requirement, fuel must be added to the mixture (Alberta Energy Regulator,

2014).

The UK has experience in dealing with enclosed flares on landfill sites. The EA has published

guidelines for monitoring such flares (Environment Agency, 2010).

There are addition technologies that allow emissions from various sources to be controlled;

these sources include (Broomfield and Donovan, 2012):

Emissions resulting from the unloading of the liquids from the borehole that are

produced after the hydraulic fracturing stage. The emissions can be controlled using

a plunger lift system which allows the surface valve to open only when the pressure

within the well has reached a certain level. This system is also a more efficient method

of removing liquids from the well.

91

Storage tanks produce emissions which can be reduced through the use of vapour

recovery units. Transfer directly to the pipeline also reduces emissions. Flaring can

be used to reduce emissions in comparison to fugitive emissions but is the least

desirable method.

Glycol dehydration units (used to separate gas from water in flowback fluid) produce

fugitive emissions, the amount of which can be reduced through the use of vapour

recovery units. Desiccant dehydrators (more efficient versions of glycol dehydration

systems) further reduce fugitive emissions (Natural Gas STAR Partners, 2006c) and

flash tank separators (incorporated into the glycol dehydration system) improve the

methane capture rate of the system (Natural Gas STAR Partners, 2006a).

Emissions from pneumatic controllers, i.e., pressure regulators and valve controllers

can be minimized by proper and rigorous maintenance. The bleed rate of a pneumatic

controller is the rate at which it emits natural gas, this is considered to be a normal

operating procedure (Natural Gas STAR Partners, 2006b). Low bleed technologies can

also be used to reduce emissions. The bleed rate can be lowered through the

application of a number of techniques. Firstly, the controllers can be replaced with

low bleed controllers that otherwise provide similar performance; secondly, existing

high bleed controllers can be retrofitted with low bleed device (Natural Gas STAR

Partners, 2006b).

A significant environmental risk to the health of the local population is the increase in air

pollution arising from the increased amount of road traffic, transporting materials to and from

the site, and on-site power generators. In terms of traffic, the majority of the vehicles will be

trucks transporting water to the site and waste products from the site. The impact these

vehicles have will depend on the location of the site and the distance to the water source and

waste treatment plants (Department of Energy and Climate Change, 2014a). Increasing the

amount of fracturing fluid recycled on-site could also help in reducing the traffic flow to and

from the site. To lower the potential impact and improve efficiency, the engines in the

machinery and vehicles could run on three-way catalytic converters or electricity

(Department of Energy and Climate Change, 2014a).

In order to mitigate the risk posed to the local population by air pollution, the operator must

monitor the emissions during the operation and report the results to the local EA or the HSE.

The operators also have to demonstrate that the drilling and fracturing process has not

increased local air pollution to levels higher than those outlined in the initial environmental

permits (Department of Energy and Climate Change, 2014a).

If it becomes economically viable to extract shale gas in the UK, the amount of infrastructure

needed at the operation site will increase. An increase in on-site machinery, e.g., compressors

and pumps, will increase emissions unless running on electric engines or three-way catalytic

converters, while the increase in gas storage and processing equipment will increase the

92

chance of a leak taking place. Joints in the pipes along which the flowback travels can result

in small leaks if they are not well maintained. If leaks occur at multiple joints, the overall

methane leakage could be substantial. The environmental risk posed by the increase in

machinery can be mitigated by ensuring that the equipment is rigorously maintained to

industry standards (Department of Energy and Climate Change, 2014a). Additionally, if the

extracted gas is stored on-site, vapour recovery units can be used to minimise the amount of

gas unintentionally vented from the storage tanks.

Evidence of emissions

Allen et al. (2013) carried out a study of 190 onshore natural gas production wells in the US.

These wells consisted of 150 production wells, 27 well completion flowback sites, 9 well

unloading sites and 4 work over sites. The measurement of methane emissions was made

using different methods, summarized in Table 9. The methods can broadly be divided into

two categories; direct measurements and mobile downwind sampling.

The study found that different extraction processes produce different quantities of emissions.

The largest source of emissions was the pneumatic controllers used on pumps, where the

amount of methane leaked ranges from 186-396 Gg methane/yr with an average of 291 Gg

methane/yr. The smallest source of methane was from completion of fracturing and flowback

which produced an average of 18 Gg methane/yr, the range of measurements was 5-27 Gg

methane/yr. The authors noted that there exists a large range in the measurements between

sites. This has the consequence of making it difficult to draw comparisons between emissions

of the same type between sites. It also highlights the need for the emissions from each site

to be considered individually. Therefore, in order fully to determine emissions coming from

individual sites, thorough baseline monitoring would be required if exploration and extraction

were to go ahead in the UK.

93

Table 9. Table displaying the emission sources, how they are measured, the average and range of yearly emissions for 190 onshore natural gas sites in the US. (Data from Allen et al., 2013).

Emission Source Direct measurement method

Mobile downwind sampling method

Average emission per year (Gg/yr)

Range of emissions per year (Gg/yr)

Well completion Measurements taken from flowback tanks using enclosures and temporary stacks. Flow rate and composition are measured.

Downwind tracer ratio method: Release of C2H2 and N2O measured on-site. Downwind measurements of CH4/C2H2 and CH4/N2O concentration ratios.

18 5-27

Unloading Temporary stacks. Measurements of flow rate and composition made.

N/A N/A (too few measurements to give a meaningful average)

25-206

Work overs Measurements taken from flowback tanks using enclosures and temporary stacks. Flow rate and composition are measured.

N/A N/A (Limited measurements) N/A (Limited measurements)

Production leaks:-

Infrared (FLIR) camera surveys of sites and flowback rate using HiFlow device

Downwind tracer ratio method: Release of C2H2 and N2O measured on-site. Downwind measurements of CH4/C2H2 and CH4/N2O concentration ratios.

Equipment leaks

See above See above 291 186-396

Pneumatic pumps

See above See above 580 518-826

Chemical pumps See above See above 68 35-100

94

A study by Bunch et al. (2014) investigated the VOC emissions in the Barnett Shale region.

The study examined over 4.6 million data points collected from seven monitoring systems at

six different locations (Dallas and Fort Worth region) set up by the Texas Commission on

Environmental Quality. Two monitors each detected 105 VOCs, these took measurements

once every six days. The remaining four measured 46 different VOCs and took continuous

measurements. In addition, the authors examined VOC records dating back to 2000 and data

was collected up until the end of 2011. The primary VOCs of interest were benzene,

ethylbenzene, m/p-xylene, n-hexene, o-xylene and toluene. The resulting VOC levels were

compared to the federal and state health-based air comparison values (HBACVs) for air quality

with the aim of assessing the health impact of the emissions. It should be noted that the

measuring systems measured the ambient air, and as such, cannot differentiate between

sources of VOCs. Therefore, emissions from other sources, i.e., from traffic were inevitably

included in the measurements.

Of the VOCs measured, all but one did not exceed the HBACVs, they therefore concluded that

the VOCs emitted posed an acceptable chronic health risk and that the population living near

to shale gas sites are not being exposed to VOC levels that would be considered to pose a

health risk. The one VOC that did exceed its HBACV was 1, 2-dibromoethane, a chemical used

in fumigant pesticides, is not associated with shale gas operations.

The authors point out that one limitation of the study, and in air monitoring studies in general,

is the fact that these studies only give an indication of the potential exposure to VOCs. The

amount of exposure varies from person to person depending on lifestyle factors, e.g.,

whether they smoke and the proximity of their home and work place to the drilling sites. The

most effective way to address this problem is through bio-monitoring studies (Bunch et al.,

2014). Another shortcoming of this study was pointed out by Werner et al. (2015). They

indicated that the systems used for the sampling were designed for regional atmospheric

measurements, not at community level. Therefore, they lacks the resolution to define

variations in the level of hazard between communities.

McKenzie et al. (2012) carried out a similar study to Bunch et al. (2014) in which they

estimated the chronic and sub chronic non-cancer hazard indices and cancer risks in

populations living less than, or more than 0.5 miles away from drilling sites. The authors

found that those living less than 0.5 miles from sites were at greatest risk compared to those

living more than 0.5 miles from the sites. Persistent exposure to the emissions from sites

during the completion stage of the wells was found to pose the largest health risk. The

emissions which augmented the hazard were determined to be trimethyl-benzene, xylenes

and aliphatic hydrocarbons. Baseline emissions were not available during this study, thus

limiting their utility. However, the spatial relationship between air quality and shale gas sites

was found to be statistically significant, hence the results appear to be valid (Shonkoff et al.,

2014).

95

These results are in conflict with those of Bunch et al. (2014) who found that shale gas

operations has not resulted in community-wide exposure to VOCs. The difference between

the results of these two studies can be attributed to use of different measurement

techniques, the volume of data analysed, and the timescale over which the data was

collected. Whilst McKenzie et al. (2012) measured for 78 different hydrocarbons, their data

set was considerably smaller than that of Bunch et al. (2014), comprising of 163 air samples

taken from stationary monitoring systems around the perimeter of shale gas sites in addition

to 24 air samples taken at the perimeters of sites during the well completion phase. However,

unlike Bunch et al. (2014), McKenzie et al. measured the air quality closer to drilling sites. By

doing this, they were able to detect more subtle variations in local air quality that would

effectively be averaged out in regional measurements (Shonkoff et al., 2014). Therefore,

these studies highlight the difference made by taking measurements on different scales.

It was made clear during the Task Force visit to the National Physical Laboratory that one of

the most significant challenges with studies focused on monitoring of emissions at shale gas

sites in the US is that each study used different monitoring techniques. As a result, comparing

the results of different emissions studies is fraught with limitations. However, studies that

are currently in preparation have tried to address this by inviting multiple institutions to

monitor the same shale gas site using their own techniques. The results of these forthcoming

studies will further greatly the understanding of the differences between monitoring

techniques, in addition to providing some insight as to how comparable these studies are and

how comparable the techniques used.

96

Food issues associated with shale gas

There is little in the way of literature on the effects of shale gas exploration and extraction on

the human food supply. Contamination of soil and surface water (i.e., those in streams, rivers

and irrigation pools) on, or near, to agricultural land can have an impact on the food grown

there. If the land is also used for grazing by livestock, any contaminants could build up in their

tissues, resulting in a potentially greater risk to the public if these animals enter the human

food chain. Few studies addressing the issues of soil contamination as a result of

unconventional natural gas operations were found (Coons and Walker, 2008; Witter et al.,

2008b).

Werner et al. (2015) covered the risks to animals (wild, domestic and farmed) in their review

paper on the environmental impacts of unconventional natural gas operations. They found

that only a small number of studies had addressed the issue (Adam and Kelsey, 2012;

Bamberger and Oswald, 2012; Finkel and Hays, 2013; Finkel et al., 2013b). However, these

studies did not address the effect that contaminants from shale gas operations have on

animals that could potentially enter the human food chain, therefore acting as an exposure

pathway. The studies instead focused on the potential for animals to act as “sentinels” for

potential long term health effects in humans. This is on account of their commonly shorter

lifetimes and reproductive cycles.

97

Health issues associated with shale gas

Before proceeding there should be a distinction made between health issues and

quality-of-life issues both of which are of concern to the general public. Health issues may

represent a longer term risk and the impact, if any, may not become apparent for a number

of years. Quality-of-life issues, such as noise pollution, light pollution and odour problems,

are likely to have a more immediate impact and may have further knock on effects with regard

to health. For instance, increased noise pollution and light pollution may lead to a higher

occurrence of stress related illnesses.

Long term studies on the health impact of shale gas extraction in the US are currently being

carried out (e.g., Marcellus Shale Initiative Study) (New York State Department of Health,

2014). There are a number of current studies that address the issue of health risk, but owing

to the fact that large scale extraction is a relatively new occurrence, the number of studies is

small and more studies are needed in order to draw firm conclusions. This is exemplified by

the study of Hill (2013) who found that the number of children born to mothers who lived

within 2.5 km of an existing well had an increased incidence of low birth weight and babies

that were small for their gestational age compared to mothers who lived within 2.5 km of a

un-drilled well site. No statistically significant differences were found in cases of premature

birth, congenital defects and infant death. Based on these findings, it was concluded that a

relationship exists between birth weight and size at birth (i.e., whether the infant was small

for its gestational age). However, the report by the New York State Department of Health

(2014) points out that the conclusion may be overstated on account of being based on a single

study. Therefore, more studies are needed in order to evaluate the findings.

Another challenge with health studies such as this one, where the general population forms

part of the data set, is ensuring that as many variables as possible are controlled. The New

York State Department of Health report (2014) highlighted that potentially significant risk

factors were not incorporated into Hill’s study, for example the quality of prenatal care, the

lifestyle of the mother during pregnancy and any pre-existing chronic diseases. It would be

difficult for one study to control adequately all the potential variables, hence reinforcing the

need for more studies.

It is worth bearing in mind that the vast majority of current studies are based on evidence

from the US where shale gas has moved from exploration into full production and extraction.

It is therefore possible that US studies will be of limited relevance to the UK at the current

stage of shale gas exploration and development.

Shonkoff et al. (2014) described a pathway that provides a connection between natural gas

and the health effects associated with it (Figure 25). In order for the full health effect to be

fully quantified, all of the stages in Figure 25 must be accounted for.

98

Figure 25. Visualisation of the environmental exposure pathway for contaminants and pollutants to result in adverse health effects in humans. The source of the contaminants is the well pad and any associated infrastructure; these produce emissions that can potentially contaminate the air, water and soil. The concentrations of these contaminants directly influences the exposure to humans whether it be through air inhalation, contact with skin and eyes, or food and water consumption. Once the level of exposure is established, the dose can be estimated over a given length of time. This, in turn, determines the health effects. (From Shonkoff et al., 2014).

Health complaints reported by the general population living near shale gas wells are skin

rashes, skin, eye and throat irritation, nausea, vomiting, respiratory problems, nosebleeds,

stress, headaches, dizziness, muscle and joint problems and a metallic / bad taste in one’s

mouth (Bamberger and Oswald, 2012; Steinzor et al., 2012; Finkel and Hays, 2013; Rabinowitz

et al., 2015). It should be noted that Steinzor et al.(2012) used a small sample size; 108 people

from 55 households in 14 Pennsylvanian counties. Five of these counties make up 85% of the

surveys submitted meaning that the data is potentially skewed towards particular

characteristics of these counties. The report also neglected to carry out control group surveys

in areas not impacted by gas extraction. The study was not comprehensive insofar as survey

forms were sent out to households and there was no requirement to fill and return them.

One also has to consider that individuals who have had a bad experience related to shale gas

extraction may be more likely to return the surveys than those who have not been affected.

These factors could act to overstate the health impact of these studies and should be treated

with caution. These problems exemplify the limitations of many of the currently published

health studies. The New York State Department of Health report (2014) emphasizes the point

that, although the health complaints may give some indication as to the risks posed to those

living close to a well site, the currently published studies can only reliably be used to generate

hypotheses upon which more thorough future research can be based. Research should centre

around epidemiology studies involving control groups from areas free from extraction of gas

to form baselines and rigorous control of bias associated with sample location, chance

findings and temporality (New York State Department of Health, 2014).

99

Recent studies

The following sections contain detailed breakdowns of three recent major studies on the

potential health impacts of shale gas operations. Each of the reports is examined in detail.

The arguments, together with the evidence cited by the authors, are presented. The

references cited here are, for the most part, those cited in the original reports. A great deal

of evidence for the potential health effects of shale gas operations cited by the reports has

already been discussed and critically appraised in previous sections of this briefing document.

The environmental health impacts review carried out by Werner et al. (2015)

Werner et al. (2015) carried out a review of the currently published evidence for

environmental health impacts of unconventional natural gas exploration and extraction. This

study presents a very thorough review of the currently published literature, hence the

majority of information presented in the following section is drawn from this paper. Any

references that are in the following section are those cited by Werner et al. (2015).

They examined literature published between January 1995 and March 2014. Through

searching online databases of academic literature, the authors initially found 109 relevant

studies. Of these, 7 were considered to be “directly relevant” based on the strength of the

presented evidence (Texas Department of State Health Services, 2010a; Steinzor et al., 2012;

Fryzek et al., 2013; Hill, 2013; Perry, 2013; Steinzor et al., 2013; McKenzie et al., 2014). 38

were considered to be “relevant” and 64 studies were considered to be “not very relevant”.

It was highlighted that all of these studies failed adequately to address long term health

impacts, such as cancer associated with pollution from shale gas sites. The authors argue that

this is reasonable given that shale gas extraction has only taken off in the last few years;

therefore, the effects may not yet be apparent. Studies that make direct associations

between shale gas operations and health impacts were considered to lack the rigour in terms

of methods thus limiting the reliability and applicability of the results. The authors do,

however, highlight the fact that just because there is currently no concrete evidence linking

the shale gas with health problems does not mean that links do not exist and vice versa.

Therefore, more studies are needed in order to understand better what, if any, links exist.

Impact on water

Werner et al. (2015) cite particular chemical additives used in the hydraulic fracturing fluid

that may be cause for concern. Although the amount by volume of chemicals used is low

(approximately 2% or less), because the amount of fluid needed hydraulically to fracture a

well is large (5 million US gallons for several frack stages) this could result in tonnes of

chemical additives being used over the course of the fracturing stages (Finkel and Hays, 2013;

100

Finkel et al., 2013a; Goldstein and Malone, 2013). With regard to ground and surface water

contamination, studies (e.g., Gordalla et al., 2013) have suggested that flowback fluid poses

a more significant hazard than the raw hydraulic fracturing fluid, on account of the salts,

heavy metals and NORM originating from the target formation, that may be present in the

flowback. Flowback can also include polycyclic aromatic hydrocarbons (PAHs), other

hydrocarbons, NORM and salts, all of which present problems for disposal. If chemicals within

the hydraulic fracturing fluid were to enter into the surface and drinking water system they

might result in adverse health effects. This is particularly the case with endocrine-disrupting

chemicals (EDCs), of which there are more than 100 used in the extraction process (Kassotis

et al., 2014). Concentrations of as low as a few parts per billion can be expected to have

adverse effects (Colborn et al., 2011). EDCs chemicals are those that can interfere with the

body’s endocrine (hormone) system (National Institute of Environmental Health Sciences,

2015). They can produce adverse developmental, reproductive, neurological and immune

effects (National Institute of Environmental Health Sciences, 2015). It should be noted that

endocrine disruptors are found in many everyday products such as polycarbonate plastics,

food and cosmetics (National Institute of Environmental Health Sciences, 2015), and exposure

occurs through ingestion of food, dust and water; inhalation of gases and air-borne

particulate matter; and through the skin (World Health Organisation, 2015a).

Few studies have thus far attempted directly to link the chemicals used in fracturing fluids to

the health effects of exposure to specific chemicals. Colborn et al. (2011) discussed how

specific chemicals, if inhaled, ingested or absorbed through the skin, could produce adverse

health effects. Compounds such as (2-BE) ethylene glycol monobutyl ether, acetic acid,

ethylene glycol, isopropanol (propan-2-ol), methanol and sodium nitrate can cause adverse

effects to the skin, eyes, immune system, nervous system and internal organs (Kargbo et al.,

2010; Colborn et al., 2011). Werner et al. (2015) note that these studies did not address the

environmental exposure pathways, actual exposure doses and the causality in terms of health

effects that these chemicals can have on the population living near to shale gas sites. These

essential facts remain to be determined and requires further study.

Impact on health related air quality

Werner et al. (2015) found that the majority of literature published on the impact of shale gas

operations on air quality focused on emission inventories and air sampling, not on direct

health impacts, although some characterisation of the risk associated with the emissions was

carried out. It should be noted that the authors also found that there is a geographical bias

in the current literature. The vast majority of current literature comes from Garfield County,

Colorado, although there is some research from Texas and Pennsylvania. Of the published

studies, two general areas of concern were found. First, hazard descriptions of air-borne

pollutants and how these are released into the atmosphere, and secondly, the health

concerns associated with the pollutants.

101

The most commonly cited pollutant is methane from fugitive emission sources. Other

pollutants that can be released during shale gas operations have been cited as potentially

harmful; these include benzene, carbon monoxide, hydrogen sulphide, nitrogen oxides,

sulphur dioxide and VOCs (Witter et al., 2008b; Colorado Department of Public Health and

Environment, 2009; Colborn et al., 2011; Kaktins, 2011; Zielinska et al., 2011; Down et al.,

2012; Weinhold, 2012; Kibble et al., 2013).

Werner et al. (2015) found that data on inhalation-related toxicity is limited, with only a few

studies providing data that are of sufficient utility. They provide one example where a study

identified 86 contaminates, the data for 65 of which were not extensive (Colorado

Department of Public Health and Environment, 2010). Therefore, the potential health impact

could be underestimated. The authors suggest that concurrent bio-monitoring studies should

be carried out with air monitoring studies because, whilst the amount of emissions can be

monitored, the amount that is actually inhaled by members of the population is not

measured. Few examples of bio-monitoring studies were found, but one such study was

carried out by the Texas Department of State Health Services (Texas Department of State

Health Services, 2010a). They collected blood and urine samples from 28 people living in and

around the town of Dish with the aim of investigating the impact of VOCs on local

communities (Texas Department of State Health Services, 2010a). They found that VOC levels

were not consistent with those that would be expected for community wide exposure to shale

gas related VOCs and pollutants. However, the study also considered that their results could

be affected by exposure to other factors, smoking, disinfectant by-products in drinking water

and workplace exposure which were not accounted for. Only one sampling period was carried

out during the study. Therefore, changes in VOC exposure with time could not be measured

(Texas Department of State Health Services, 2010a). These factors represent a shortcoming

of the sampling protocols used in this study, and others carried out to date, and indicate the

need for appropriate sampling to be carried out in future studies. If this is done, it may be

possible to differentiate between VOC exposure originating from shale gas operations and

those from other sources in addition to determining the variation in VOC levels with time.

Pollutants in soil

Few studies addressing the issues of soil contamination as a result of unconventional natural

gas operations were found (Coons and Walker, 2008; Witter et al., 2008b).

Contamination of soils mainly occurs through spills and leaks on-site but can also occur when

the drilling cuttings are being stored, transported and disposed of (Zoback et al., 2010). Some

of the same pollutants cited as potentially dangerous when emitted from fluids, e.g. benzene,

can also contaminate land by either adsorbing into or absorbing onto soil particles. This

creates a residue that can leach upon interaction with rain or snowmelt. Pollutants can also

be inhaled, absorbed or otherwise ingested (Coons and Walker, 2008; Witter et al., 2010).

102

Werner et al. (2015) make the point that data on soil quality and contamination, or lack

thereof, in relation to shale gas operations is lacking. They also highlight the need for baseline

studies in order fully to quantify the impact of any contamination.

Occupational health

The above discussion has only addressed the issue of the hazard and risks posed to the wider

population by shale gas operations. However, the hazard and risk to the on-site employees

is of significance as they will come into contact with substances, such as flowback fluid stored

in tanks, on a daily basis. Studies have highlighted that safety hazards in the oil and gas

industry are well documented but health hazards and chemical exposure risks are less well

studied (Coussens and Martinez, 2013; Esswein et al., 2013). A major occupational health

hazard that could potentially affect on-site employees is that of crystalline silica (i.e., sand

and dust) as, when inhaled, can result in silicosis, lung cancer, tuberculosis, autoimmune

diseases and kidney disease (Laney and Weissman, 2012; Esswein et al., 2013). Silicosis is of

particular concern as it has a long latency period (as long as decades) between exposure and

the development of symptoms (Laney and Weissman, 2012). A recent exposure assessment

study of crystalline silica on shale gas sites in the US found that many of the samples taken

were above the acceptable exposure limit (Esswein et al., 2013). As would be expected,

workers using certain equipment, i.e., sand mixers, were exposed to the greatest amount of

crystalline silica (Esswein et al., 2013). Building site and quarry workers can be exposed to

very high levels of silica dust exposure.

Two studies were found that examine the potential effect of drilling fluid (mud) on the health

of on-site employees (Broni-Bediako and Amorin, 2010; Searl and Galea, 2011).

Broni-Bediako and Amorin (2010) found that exposure to fluid was mainly through inhalation

of vapour mist, dermal or oral contact with vapours, aerosols and dust. Searl and Galea (2011)

similarly found that the main health risks are inhalation of vapour and aerosols. They also

said that long term exposure can increase the risk of chronic respiratory illnesses in addition

to neurological problems potentially including dementia. It should be noted that the risks

associated with exposure to fluids associated with shale gas operations will vary from site to

site as it is highly dependent upon the composition of the fluids used.

Noise pollution was also cited as a potential health hazard by (Witter et al., 2014). Prolonged

exposure to noise emanating from compressors, generators, drilling, diesel engines,

mechanical brakes, radiator fans and heavy machinery can result in noise-related health

problems, such as deafness and tinnitus (Witter et al., 2014). This applies to many types of

industrial activity.

103

Health impacts from infrastructure associated with shale gas operation

When compared to other health issues, i.e., air and water pollution, the number of studies

published regarding the health impact of shale gas related infrastructure is much lower.

A study by La Plata County (2002) considers that, although the drilling sites may be situated a

sufficient distance apart as to minimise the amount of noise pollution coming from a single

site, the cumulative effect of multiple sites may be larger depending on each site’s stage of

operation. Low frequency (infrasonic) noise pollution, i.e., that emanating from the

continuous running of generators, has the potential to cause various health problems, i.e.,

stress, annoyance, irritation, fatigue, headaches, unease, disturbed sleep and cardio vascular

problems (Witter et al., 2008a; Witter et al., 2010; Witter et al., 2013).

Werner et al. (2015) found that no studies had investigated the impact of light pollution

resulting from shale gas operations, whether on humans or wildlife. Some studies have

suggested measures, such as directional lighting, modified drilling rig placements and glare

restrictions, that can reduce the potential for light pollution (Witter et al., 2010; New York

State Department of Environmental Conservation, 2011), however, no studies have gone as

far as to assess the risk of the light pollution hazard. This may be significant as sites will likely

be running 24 hours a day. Recent studies from other industries has suggested a link between

artificial light exposure and increased cancer risk (Witter et al., 2008a). This of particular

concern to those employees working night shifts, as they will have prolonged exposure to

artificial light. Again, this hazard applies in many spheres of employment.

Increases in traffic flow may occur around areas of shale gas operations but this has not been

subject to any in-depth studies from a health perspective, as opposed to road safety. Any

increase in traffic flow will likely be site specific and be influenced by the existing

infrastructure, i.e., the number of wells being drilled and the number of other sites located

nearby. In addition, the traffic flow will vary depending upon the drilling stage. Gas and

particulate emissions from heavy goods vehicles transporting material to and from sites has

been considered to result in increased air pollution around shale gas sites (Hill, 2013). UK

planning authorities refusing permits have several times appealed to the unacceptability of

increased traffic as grounds for refusal, so it is important that this perceived hazard is studied

properly and objectively.

Social impacts

Werner et al. (2015) identified three health related social impacts related to unconventional

natural gas operations; symptomatological, risk perception, and governance and regulation.

In terms of symptomatological studies, surveys of residents living in near to shale gas

operations in the US tend to feel that their health is being adversely affected (Steinzor et al.,

2012; Steinzor et al., 2013). The members of the local population that responded to these

104

surveys conveyed concerns about what they perceived to be health symptoms directly related

to natural gas operations. These symptoms were nose, eye, and throat irritation, respiratory

symptoms, nausea, nosebleeds, sleep disturbance, rash, headaches, ringing in ears,

abdominal pain or cramping, extreme drowsiness, fatigue, and weakness (Subra, 2009, 2010;

Texas Department of State Health Services, 2010a; Bamberger and Oswald, 2012; Steinzor et

al., 2012; Saberi, 2013; Steinzor et al., 2013). However, Werner et al. (2015) found no

evidence of direct cause and effect. Adgate et al. (2014) consider that many of these non-

specific symptoms could be attributed to psychological stress associated with having a nearby

unconventional natural gas site. They also consider that this could, itself, be seen as a health

risk but has not yet been studied as one.

As would be expected, the number of reported symptoms associated with shale gas

operations decreases with increasing distance from the operation site (Steinzor et al., 2012).

This could either be an indication that there is a direct link between shale gas operations and

health effects or it could indicate the psychological impact of living near to a site.

One of the most commonly reported concerns is that of odour, with local residents reporting

a range of odours, i.e., unidentified gas, sulphur, burnt butter, propane, sickly-sweet smells

and ‘chemical-like’ smells (Subra, 2009, 2010; Steinzor et al., 2013).

Cancer incidence, as a result of benzene emissions from shale gas sites, was also cited as a

social concern by communities in Texas. In a series of studies, the Texas Department of State

Health Services (Texas Department of State Health Services, 2010a, b, 2011), the cancer

incidence was investigated. The results from the most recent study found no correlation

between shale gas operations and the incidence of various types of cancer. Specifically

childhood leukaemia, childhood brain / central nervous system cancers, all-age leukaemia,

and all-all non-Hodgkin’s lymphoma, in the surrounding communities were in the expected

ranges for both sexes (Texas Department of State Health Services, 2011). However, the

incidence of breast cancer was statistically significantly higher than expected. This was

attributed to the rapid increase in population around areas of shale gas operations (Texas

Department of State Health Services, 2011). The study also considered that the number of

cancer cases upon which the study was based was likely to have been underestimated

because the population data was obtained from the 2000 Census (Texas Department of State

Health Services, 2011).

Another study, by Fryzek et al. (2013), investigated the incidence of childhood cancers at

different stages in the drilling process. They found that the occurrence of all but one cancer

type was near to expected levels both before and after drilling. There was found to be a slight

increase in the standardised incidence ratio for central nervous system tumours after drilling

had taken place. Overall, the authors concluded that there was no increase in the incidence

of cancer in communities living near to drilling sites. Goldstein and Malone (2013) disputed

this conclusion as being ‘unfounded’. Werner et al. (2015) pointed out that the timeframe

for the studies (1995-2009) does not encompass the complete development of shale gas in

105

the Marcellus region therefore the results and conclusions drawn by Fryzek et al. (2013) are

of limited applicability.

Many generalised symptoms are difficult to attribute directly to shale gas developments,

however, members of the local population have reported their symptoms becoming less

prominent when they leave the area around shale gas sites (Steinzor et al., 2013). Current

epidemiological studies (Texas Department of State Health Services, 2010b, 2011; Fryzek et

al., 2013) have not linked cancer incidence to shale gas operations. There is, however, still a

need to expand the number of studies and to conduct more research into long term

epidemiology.

Werner et al. (2015) found that the public’s perception of hazard must be considered as a

factor. Perry (2012) found that the public create their own perception of hazard and risk

based on the available scientific information and regulatory information. In turn, the

uncertainty associated with perceived risk can cause health related complaints, i.e., stress and

anxiety (Perry, 2013). Werner et al. (2015) concluded that there is a real need for studies

comparing the public perception of risk to actual risk data.

Public Health England 2013 Report

Public Health England (PHE), an executive agency of the Department of Health, carried out a

review of the potential public health impacts of exposures to chemicals and radioactive

pollutants resulting from shale gas operations (Kibble et al., 2013). The review was carried

out, specifically, by the PHE Centre for Radiation, Chemical and Environmental Hazards (CRCE)

who examined literature and data from countries where commercial-scale shale gas

extraction operations were already underway.

The review highlighted that potential risks to human health cited in literature, e.g., fugitive

emissions and surface spills of hydraulic fracturing fluid, were a result of either poor

regulation or an operational failure. The difference in potential human impact between single

well exploratory sites and full-scale extraction involving multiple wells, the cumulative impact

of which may be more considerable, was also emphasised. The overarching conclusion drawn

from the study is that the potential risks to public health from emissions will be minimal if the

operations are run and regulated properly.

Aside from the main conclusion, the review made eight recommendations:-

1. PHE should continue to work with regulators to ensure that all aspects of shale gas

exploration and extraction in order to ensure that all risks are appropriately assessed.

2. Baseline environmental monitoring is required in order to assess the impact of shale

gas operations on environmental, and hence public, health. The development of

emission inventories should be considered as part of the regulatory regime.

106

3. Effective monitoring is required throughout the entire lifetime of the site including the

post-production abandonment phase.

4. The broader socioeconomic impacts, such as increased traffic and the impact on local

infrastructure should be considered during the planning stage of any operation.

5. The chemicals that are used in the hydraulic fracturing fluid should be publicly

disclosed and assessed prior to use. The review notes that any potential risk to the

public posed by the chemicals will be dependent upon the exposure pathway,

together with the total volume, concentration and fate of the chemicals. It is

considered by the authors that these risks will be assessed during the regulatory

environmental permitting process.

6. The type and composition of gas will vary on a site by site basis, and as such, the risk

assessment should be carried out on a site-by-site basis.

7. Evidence from the US indicates that well integrity, appropriate storage and

management of hydraulic fracturing fluid and waste products are key to ensuring that

risks are minimised, therefore, the appropriate regulatory control will need to be put

in place. Again, implementation details will likely vary from site to site.

8. Characterisation of potentially mobilised natural contaminants, i.e., NORM and

dissolved minerals originating from the target formation, is needed.

Air quality

The PHE review said that, at the time of publication, there was no published data regarding

emissions from UK shale gas sites, nor has there been any emission data published from the

hydraulically fractured, conventional tight gas sand well at Elswick, Lancashire, where

production has been carried out since 1993. At this particular site, gas was extracted from

low permeability sandstone via hydraulic fracturing of a vertical well. Note that the degree

of fracturing required was far less than that required to stimulate gas production from shale.

The PHE review considered that emissions from single sites will likely be small, intermittent

and not unique to shale gas operations, i.e., comparable to certain other industries. However,

the main environmental hazard comes when the density of drilling pads and wells per pad

increases, producing potentially a larger cumulative impact (Kibble et al., 2013). Common

compounds that might be emitted from shale gas sites (i.e., nitrogen oxides (NOx) VOCs and

particulate matter) can produce secondary pollutants, e.g., ozone (O3). However, these

pollutants can also be produced from other industry sources and transport (Kibble et al.,

2013). This reinforces the need to establish background emission levels prior to any drilling

taking place.

In terms of published evidence, the review cites a study by Zielinska et al. (2011) in which air

pollutants from Barnett Shale sites were characterised. The authors found that, in addition

to methane, 70 different VOCs were found, the most abundant of which were ethane,

propane, n-butane and pentane. These VOCs made up approximately 90% of the total

107

emissions. The remaining 10% was made up by 2- and 3-methylpentane, n-hexane, methyl-

cyclopentane, cyclohexane, 2-methylhexane, 1-heptene, methylcyclohexane, n-heptane, and

n-octane (Zielinska et al., 2011). PAH (benzene, toluene and xylenes) account for 0.1-0.2% of

non-methane VOCs (Zielinska et al., 2011). The main source of the emissions was found to

be from malfunctioning condensation tanks.

Between 2008 and 2010, Rich et al. (2014) carried out emissions work in the Dallas and Fort

Worth areas overlying the Barnett Shale. The aim of the study was to attempt to establish a

“fingerprint” of chemicals that could be associated with shale gas operations in residential

areas. To do this, the authors collected ambient air samples from residential areas within

61 m of shale gas extraction / production. A total of 50 sampling data sets were obtained and

analysed. Most areas had methane levels (a mean of 11.99 ppmv (parts per million by volume)

and a median of 2.7 ppmv) higher than background urban concentrations (1.8-2.0 ppmv).

Other chemical components were found to correlate with the presence of methane, for

instance 3-methylhexane. The researchers found that seven chemicals (o-xylene,

ethylbenzene, 1,2,4-trimethylbenzene, m- and p-xylene, 1,3,5-trimethylbenzene, toluene and

benzene) could potentially provide a pollution signature for shale gas operations. However,

there are limitations to the study which are pointed out by the authors. These include the

small sample size and that correlation with a common source does not mean other sources

are not contributing to the correlation, i.e., other sources of chemicals and methane could

not be distinguished from those from shale gas operations. Rich et al. (2014) suggest that a

further study with a larger sample size and more rigorous analysis be carried out.

Roy et al. (2014) developed an emissions inventory for the development, production and

processing stages of shale gas operations in the Marcellus Shale for 2009 and projected 2020.

Based on the 2009 estimates, operations in the Marcellus Shale may account for 6-18% (12%

average which equates to 129 tons per day) of the NOx emissions and between 7 and 28%

(12% average which equates to 100 tons per day) of the anthropogenic VOC emissions, with

an average contribution of 12% (100 tons per day) in the Marcellus region. The study also

examined particulate matter. The authors consider that shale gas operations will not make a

significant contribution to particulate matter emissions in the Marcellus area. It should be

noted that the 2020 estimates are subject to considerable uncertainties due to the fact that

assumptions were made regarding future control measures on emissions (Kibble et al., 2013),

i.e., if more stringent regulations are introduced, emissions will fall. Certain unknown

parameters also introduce uncertainties related to the emission estimates. For instance, the

engine on-time of the drilling rigs is considered to cause uncertainties in NOx estimates (Roy

et al., 2014).

Using daily air samples collected at the National Oceanic and Atmospheric Administration

Boulder Atmospheric Observatory, Weld County, Colorado, together with a road based air

sampling survey, Pétron et al. (2012) examined the emissions from shale gas operations in

the Denver-Julesberg Fossil Fuel Basin, Colorado. Air samples were analysed for methane and

108

non-methane VOCs (specifically propane, n-butane, i-pentane, n-pentane and benzene). A

range of emissions producing activities took place in the basin, including oil and gas

operations (both conventional and un-conventional), a landfill site, feedlots, a water

treatment plant and automobiles. After being corrected for wind direction, the results

suggested that strong alkane and benzene signatures were coming from the north-eastern

Colorado. The main activity that produced such compounds was oil and gas operations. The

authors considered that flashing of gas from condensate tanks and venting and leaking of

wells as being key emission sources.

The type of emissions produced by each site will vary depending upon whether the gas is wet

or dry. Wet gas is produced when pressure/temperature conditions of burial were not so high

as to preclude the formation of liquid hydrocarbons. Thus wet gas contains other

hydrocarbons, e.g., butane and ethane, while dry gas does not. These other hydrocarbons

are known as natural gas liquids (Kibble et al., 2013). As wet gas contains other hydrocarbons,

there is the potential for these sites to produce more VOC emissions. Evidence from the US

(Roy et al., 2014) has shown that the gas composition can vary within a single shale formation.

The Marcellus shale is known to be mainly dry gas but in southwest Pennsylvania, wet gas is

found. This highlights the need to asses each operation site on a case-by-case basis.

In summarizing the impact of shale gas operations on air quality, the review highlights the

fact that any emissions are highly dependent upon the phase of well development,

operational practices, geology, local topography, meteorology, the type of activities and on-

site equipment. Therefore, the type and volume of emissions will be unique to each site

whilst also varying with time. The review considers that, because of these variations, it is

impossible to directly apply the findings from the US to UK shale gas operations (Kibble et al.,

2013). However, the information from the US does provide some indication of the emission

sources and how to go about managing them.

In terms of further work, PHE suggest that, because the type of gas produced will vary from

site to site, detailed risk assessments will need to be carried out at each shale gas site. Air

quality monitoring should be carried out both before and during any operation. Of particular

importance will be the need to carry out regional scale monitoring due to the potentially

significant cumulative effect of numerous well pads and wells.

Radon

Radon-222 is a product of the uranium-238 decay chain with radium-226 being the immediate

precursor. It has a radioactive half-life of 3.8 days and is released from most rocks and soils,

although the amount that is released is dependent upon mineralogy. Radon emits alpha

particles which are harmless unless ingested by breathing or via groundwater. Radon

migrates through both fractures in rocks and pore spaces in soils; the rate of migration is

controlled by the transmission characteristics of rocks, soils and underground fluids.

109

Exposure to natural radon sources accounts for 84% of the annual radiation exposure

experienced by the members of the general population (Watson et al., 2005). Shale gas wells

have the potential to act as conduits for radon migration to the ground surface if the well

passes through radon-containing formations (Kibble et al., 2013).

Radon concentrations in the open air are generally low (5-15 Bq/m3 (World Health

Organisation, 2015b)); however, where radon can be drawn into buildings, on account of

pressure gradients, the concentration is can be much higher (thousands of Bq/m3) (Kibble et

al., 2013). Therefore, indoor radon exposure poses the greatest hazard to the general

population. In the UK, the typical annual exposure to radon associated with natural gas is

4 µSv/yr (based on typical natural gas usage and a concentration of 200 Bq/m3) (Dixon, 2001).

This is smaller than the levels of radiation from other every day sources, for instance, the

typical yearly exposure is less than that of a chest X-ray (100 µSv) and smaller than the hourly

dose rate of travelling in a plane at a cruising altitude of 12 km (5 µSv/hour) (STUK - Radiation

and Nuclear Safety Authority, 2012). Note that the average UK radon exposure is

1,300 µSv/yr.

The literature review carried out by Kibble et al. (2013) found no UK specific information on

radon associated with shale gas operations. There is, however, information from the US

relating to radon in natural gas. Kibble et al. (2013) cite studies by Johnson (1973) and

Resnikoff (2012) who proposed values of radon associated with natural gas operations of

1,370 Bq/m3 and 1,365-95,300 Bq/m3 respectively. The latter figures proposed by Resnikoff

(2012) are obtained for estimated radon concentrations at the wellhead. The same study also

says that, depending on gas treatment, processing and transport length, the radon

concentration in the gas delivered to customers can be as high as 72,000 Bq/m3, resulting in

indoor radon concentrations of 20 Bq/m3. Kibble et al. (2013) point out that Resnikoff (2012)

did not account for mixing of natural gas from shale operations with gas from other sources

which could influence the radon content.

A study by Rowan and Kraemer (2012) found radon concentrations of 37-2,923 Bq/m3

between the wellhead and water-gas separator at 11 drilling sites in the Marcellus Shale. By

applying the values of Dixon (2001) and the upper concentration values of radon found at the

wellhead by Rowan and Kraemer (2012), Kibble et al. (2013) determined that exposure to

members of the public in the UK would be 60 µSv/yr. This is approximately 0.5% of the annual

radon exposure of 1,300 µSv/yr. Short-term outdoor radon concentrations of up to

165 Bq/m3 were measured around a shale gas site in Colorado by Burkhart et al. (2013). The

authors note that this is higher than normal outdoor levels (5-15 Bq/m3 (World Health

Organisation, 2015b)); however, they were unable to differentiate between sources.

In terms of radon in water, ingestion and degassing resulting in airborne radon are cited as

the most significant sources of radon exposure (Kibble et al., 2013). Kibble et al. (2013) say

that the largest radon concentrations are found when groundwater comes into contact with

crystalline rocks (igneous / metamorphic rocks). A study by Otton (1992) is cited as evidence

110

of increased radon concentration around hydraulic fracturing wells; although Kibble et al.

(2013) note that the cause of the increased radon concentration was not established, it could,

therefore be a natural occurrence. Assuming migration of radon from shale formations into

groundwater as a result of hydraulic fracturing, Myers (2012) determined that the time

required for migration via advection would result in any radon decaying to the point where it

would no longer pose a threat to humans. Overall, Kibble et al. (2013) consider that, due to

the short half-life of radon and the depth at which hydraulic fracturing will take place in the

UK, any radon that enters into the groundwater system would have minimal impact.

Radon, because of its solubility, can be present in flowback fluid. Kibble et al. (2013) found

no information directly relating to the concentration of radon in flowback fluid.

Measurements of alpha particles, which are emitted, together with beta particles, as radon

decays, in the Preese Hall flowback water have been made by the EA (Environment Agency,

2011). Gross alpha particles were found in concentrations of 10-200 Bq/L; radium-226, the

pre-cursor to radon, was found in the fluid. This is considered to indicate the probable

presence of radon (Kibble et al., 2013). Note that, as pointed out previously, alpha particles

need to be ingested in order to pose a hazard. Degassing of radon at the wellhead is

considered by Kibble et al. (2013) to have the potential to result in localized increases in

radon. As such, the authors consider it to be an occupational hazard, for instance, to workers

in the unlikely event of eating their sandwiches at the wellhead, as opposed to a public health

hazard.

Kibble et al. (2013) conclude that radon could present a highly localized hazard on-site but is

not likely to lead to a significant public health hazard.

NORM

NORM are considered by Kibble et al. (2013) to be present in cutting fluids, drilling muds and

flowback fluids. NORM concentrations have been measured by the EA in the flowback fluids

from the Preese Hall site and found levels to be similar to that found in granite (Environment

Agency, 2011). The fluid contained high levels of sodium, chloride, bromide and iron

(Environment Agency, 2011). Levels of lead, magnesium and zinc were higher than those in

the mains water used in the initial fracturing fluid (Environment Agency, 2011). Naturally

occurring radionuclides of potassium-40, lead-212, lead-214, bismuth-214, radium-226 and

actinium-228 were found. The highest concentrations in fluid were associated with radium-

226 (14-90 Bq/L) while the radium-226 activity of suspended solids was found to be 2.5-

7.2 Bq/kg. Kibble et al. (2013) say that these values are consistent with the range of values

found elsewhere in Europe (0-200 Bq/L and 5-900 Bq/kg).

With regard to assessing the health risks, Kibble et al. (2013) emphasise the challenges posed

by accurately determining the exposure of the general public to NORM. Specifically, they

describe the need to have detailed knowledge of the operation, the everyday habits of both

111

the workers and local population, and information about the treatment and disposal of any

radioactive material. There is also a need for differentiation between the hazards posed by

drilling waste and those posed by flowback fluid.

Kibble et al. (2013) consider that the UK has a history of dealing with NORM contaminated

drilling waste that has been produced from traditional hydrocarbon operations. As such,

there are procedures and waste management plans in place to ensure that waste is properly

handled and disposed of. A potentially more challenging prospect is that of NORM

contaminated flowback fluid. The authors point out that the waste treatment plants will have

specifically to address factors including the storage, treatment, transport and disposal of the

fluid.

(Kibble et al., 2013) are also of the opinion that, based on current measurements from the

Preese Hall site together with measurements and assessments of the radioactivity exposure

experienced by offshore oil and gas workers, it is not expected that shale gas operations

would pose a radiological hazard to the public. They also consider that the current regulatory

system is capable of protecting both the general public and workers from the radioactivity

hazards arising from NORM. Despite this, the authors highlight the need for data on UK

NORM levels as this is the only way accurately to predict potential exposure in the UK.

Water and wastewater

The PHE report highlights the importance of ensuring that the water needed for the hydraulic

fracturing process is extracted from a sustainable source (Kibble et al., 2013). Potential

problems associated with the storage of the hydraulic fracturing fluid are also raised. The

authors say that, if fracturing fluid is stored on site, there is an increased chance of

contamination of surface waters if a spill takes place due to poor handling or

mismanagement. In addition, it is suggested that, because of the volume fluid returning to

the surface as flowback, sites may not have the processing capacity to deal with all the waste.

Although the authors acknowledge the potential for recycling of fracturing fluid, they also say

that the volume that can be recycled is dependent upon the salinity and concentration of

chemical constituents accumulated during the fracturing and flowback process.

Kibble et al. (2013) say that there is currently no peer reviewed literature on the impacts of

shale gas operations on water in the UK or elsewhere in Europe. However, there have been

some non-peer reviewed studies that have looked at the potential impact on drinking water

sources (Broderick et al., 2011; House of Commons, 2011; The Royal Society and the Royal

Academy of Engineering, 2012). As has been discussed in the water contamination section,

of this briefing document, contamination of ground and surface water has taken place in the

US as a result of shale gas operations. The Royal Society and the Royal Academy of

Engineering report (2012) also highlighted the need for baseline methane measurements in

groundwater prior to any operations being undertaken. The BGS are currently in the process

112

of establishing these baselines. The analysis of the Preese Hall flowback fluid showed high

levels of sodium, chloride, bromide and iron together with levels of lead, magnesium, zinc,

chromium and arsenic elevated above the concentration found in the water used to form the

basis of the fracturing fluid (Kibble et al., 2013).

Kibble et al. (2013) cite studies discussed previously as evidence of water contamination

resulting from shale gas operations (Osborn et al., 2011; Warner et al., 2012; Jackson et al.,

2013; Molofsky et al., 2013; Vidic et al., 2013; Warner et al., 2013). The authors also discuss

a study by the Massachusetts Institute of Technology (2011) which reviewed 43 incidences of

water pollution that took place between 2005-2009. The study found that loss of well

integrity is the major cause of contamination (20 of the 43 incidents) with surface spills being

the second most important cause (14 of 43 incidents). Contamination took place via

inadequate cementing, inadequate casing installation, hose leaks at the surface, overflowing

pits and failure of the pit lining. As has been noted previously, the use of open pit storage is

prohibited in the UK; therefore the potential for contamination events from these last two

sources will be non-existent. A study by Goss et al. (2013) which analysed groundwater

benzene, toluene, ethylbenzene and xylene levels after surface spills had taken place in the

US is also cited in the PHE report. The study found that levels of these contaminants exceeded

US the safe drinking water limits in some samples, however, the remedial action taken by the

operators reduced these levels effectively.

The PHE report addresses concerns regarding drinking water contamination during hydraulic

fracturing by saying that 99% of water that enters people homes comes from water

companies that have not only carried out treatment on the water, but also have a number of

additional quality control measures in place (Kibble et al., 2013). These include

measurements of water quality as it leaves the treatment facility in addition to random

samples taken from homes. Kibble et al. (2013) see it as unlikely that contaminated water

would enter into homes on account of this regulatory scheme. The authors also consider

contamination as a direct result of hydraulic fracturing, i.e., upwards migration of fluids, to

be a minimal risk owing to a lack of current evidence that suggests that fractures can migrate

a sufficient distance to interact with groundwater.

Citing a publication by the Chartered Institute of Water and Environmental Management

(2014), Kibble et al. (2013) point out the importance of considering the following when

planning a shale gas operation; baseline monitoring; monitoring throughout the lifecycle of

the well; the potential contamination resulting from a loss of well integrity; the need to follow

best operational practices; the transport, management and storage of chemicals, hydraulic

fracturing fluid and flowback fluids; and the treatment of flowback fluids. The Chartered

Institute of Water and Environmental Management (2014) suggest hydraulic fracturing

should not be allowed to take place in areas where there is a risk to groundwater. In addition,

it is considered that the relationship between groundwater and the target shale formations

113

be examined before any drilling takes place (Chartered Institute of Water and Environmental

Management, 2014).

Kibble et al. (2013) note that Water UK, the representatives of the UK water industry, and

UKOOG have signed a memorandum of understanding that both bodies will work together

to minimise the environmental impact of any future shale gas operations. This approach

includes baseline monitoring, water management and waste water disposal (United Kingdom

Onshore Operators Group, 2013b).

Kibble et al. (2013) highlight the potential environmental impact of waste waters and

flowback fluid once they have been treated at the appropriate waste management facilities.

The authors cite the study carried out by Warner et al. (2013) which has been previously

discussed in the surface water contamination section. In addition, studies by Voltz et al.

(2011), Hladik et al. (2014) and Shariq (2013) are also cited. The study by Voltz et al. (2011)

found that contamination of tributaries of the Ohio River by barium, strontium and bromide

had taken place after treated waste water from shale gas operations had been released. The

authors also suggested that the increased bromide concentration could lead to the formation

of disinfectant by-products (DBP) in chlorinated drinking water. The same suggestion was

made by Hladik et al. (2014) who found, from examination of waste water treatment plants

in Pennsylvania that accepted waste from both conventional and unconventional

hydrocarbon operations, that the level of DBPs was increased compared to treatment plants

not accepting waste from hydrocarbon operations. Shariq (2013) raised the issue of treated

and diluted wastewater being used to irrigate crops. Kibble et al. (2013) say that, although

there is little data on the issue, the practice is unlikely to be permitted in the UK.

Hydraulic fracturing fluid

Kibble et al. (2013) found no peer reviewed literature relating to the composition and use of

hydraulic fracturing fluid in the UK. This is partly down to the lack of drilling activity but also

to the need to consider each site individually. The composition of the hydraulic fracturing

fluid that has been used at the Preese Hall site can be found in Appendix 2.

Kibble et al. (2013) cite studies by Kassotis et al. (2014) and Colborn et al. (2011), the latter of

which have been discussed as part of the Werner et al. (2015) study, as presenting evidence

of the potential effect of the chemicals used in the hydraulic fracturing process. The study by

Kassotis et al. (2014) was motivated by the authors estimate that there are potentially 750

chemicals and components that can be used in the hydraulic fracturing and gas extraction

processes. More than 100 of these chemicals and components are known or are suspected

to be endocrine disruptors.

Kassotis et al. (2014) investigated 12 EDCs used in hydraulic fracturing fluids in the Garfield

County, Colorado. 39 unique surface and ground water samples were taken at 8 sites

including 2 reference sites and the main drainage basin for the sample sites (Colorado River).

114

These sample collection sites were those that had experienced some form of surface spill or

leak. Samples collected from drilling-dense areas showed higher concentrations of these

EDCs compared to reference sites where limited, or no, drilling took place. This suggests that

baseline monitoring is needed before drilling starts in order better to establish pre-existing

levels of EDCs. Samples from the Colorado River displayed moderate amounts of EDCs, larger

than the control sites but lower than the spill / leak sites. The study considered that spills and

leaks result in higher concentrations of EDCs in surface and ground water. It should be noted

that Kassotis et al. (2014) do not categorically state that water contamination took place due

to drilling activity, but rather they suggest that increased EDCs may be a result of drilling

activity. In addition, both naturally occurring chemicals and those originating from drilling

activity were measured in this study, although the contribution from natural sources was

considered to be minimal. The study also gives no indication given regarding the volume of

leaked fluid. Kassotis et al. (2014) concluded that shale gas activity may result in an increase

in endocrine disrupting compounds in groundwater. Kibble et al. (2013) point out that the

study was carried out using an in-vitro (cell culture) test system and that endocrine disrupting

activity found in such test systems may not necessarily translate into endocrine disrupting

activity in in vivo systems (i.e., entire organisms). They can, however, prove to be useful in

informing decisions regarding further toxicity testing in addition to indicating potential modes

of action (i.e., changes at a cellular level) (Kibble et al., 2013).

The PHE report considered that, although the volume of chemicals used in the process of

hydraulic fracturing is small, the potential impact of multiple wells could be considerable. The

prospect of these chemicals being stored on-site could also prove to be hazardous. Because

of this, Kibble et al. (2013) emphasise the need for transparency in disclosing the chemicals

in fracturing fluids and the need for such chemicals to be subjected to independent

assessment before use. This will allow a thorough hazard and risk assessment to be carried

out. This is particularly significant as the composition of fracturing fluid will likely vary from

site to site. Suitable accident management plans and enforcement of best practices and

regulations are also considered by Kibble et al. (2013) to be key to minimising the impact of

any accidents that might occur on-site.

New York State Department of Health 2014 report

A similar public health review was carried out by the New York State Department of Health

(2014). Like the PHE study, the New York State report covered a range of potential areas that

could impact on human health, specifically, air pollution, climate change impacts, drinking

water impacts, soil and water contamination, inadequate treatment of wastewater,

earthquakes and socioeconomic impacts.

The review determined that the impact on public health is difficult to assess fully as the

hazards and associated risks will vary spatially on account of differing well pad densities,

populations and baseline environmental conditions. The authors consider that, because of

115

the dispersed nature of the potential operation locations in New York State, the chance of

equipment failure and / or process failures is increased (New York State Department of

Health, 2014). This, in turn, could lead to an increased cumulative risk of exposure. However,

if appropriate regulations were in place and adhered to, one would assume that the hazards

associated with equipment and process failures should be reduced although a guarantee of

absolute safety cannot be made (New York State Department of Health, 2014).

The New York State Department of Health (2014) review highlighted the lack of long term

studies on the health impacts of shale gas exploration and extraction. At the time of

publication, these long term studies had either not been published or had not been started

(New York State Department of Health, 2014). The available information was considered to

be only exploratory in nature and demonstrates considerable uncertainties in the human

health impact of shale gas operations (New York State Department of Health, 2014).

The review cited a number of long term studies currently underway in the US. The Marcellus

Shale Initiative Study, which began pilot studies in 2013, is one of the long term studies being

carried out. However, the results will not be available for many years (New York State

Department of Health, 2014). The study aims to assess the impact of shale gas operations on

30,000 asthma patients and 22,000 pregnancies between the years 2006-2013, through the

use of exposure estimates. The University of Colorado at Boulder, working in conjunction

with the National Science Foundation (NSF), are in the process of carrying out a number of

investigations into assessing and mitigating problems posed by shale gas operations. The co-

operative is set to extend to 2017 with research being published throughout its lifetime. The

EPA are currently carrying out a study on the potential impact of hydraulic fracturing on

drinking water resources and to establish the driving forces that determine the severity and

frequency of contamination events. The study began in 2011 and the complete results are

not expected to be published until 2016. The New York State review suggests that the results

of these studies will reduce the uncertainty associated with assessing the risks and

environmental impact of shale gas operations, however, the results may not become available

for a number of years.

The Pennsylvania Department of Environmental Protection have carried out a Comprehensive

Oil and Gas Development Study. The study, which began in 2013, is analysing the

concentration of radioactive components in flowback waters and waste residues produced

from shale gas operations together with radon measurements in the natural gas. In addition,

the potential exposure of the public and site workers is being investigated (New York State

Department of Health, 2014). Results from this study have recently been published (Perma-

Fix Environmental Services Inc., 2015). The study concluded that natural gas extraction poses

little threat of increased radon exposure to the public (Perma-Fix Environmental Services Inc.,

2015), a stance similar to that of the PHE report. It was also concluded that there is little or

limited potential for radiation exposure to either the general public and workers, either

on-site or at any stage in the production and supply chain (Perma-Fix Environmental Services

116

Inc., 2015). The main environmental radiological hazard is that of fluid spills. Wastewater

treatment plants should be monitored for elevated levels of radioactive material and, if

found, radiological discharge limitations and spill policies should be considered (Perma-Fix

Environmental Services Inc., 2015). It was considered that landfill sites that receive treated

waste products present little potential for increased exposure to radioactive material,

although it was proposed that the filter cake produced during the treatment of waste

products could present an environmental hazard if spilled (Perma-Fix Environmental Services

Inc., 2015). However, as the main exposure pathway is ingestion, the filter cake will have to

infiltrate the drinking water supply in order for it to pose a risk to humans. The disposal of

the filter cake could also present a long term environmental hazard. The study suggested that

the protocols governing the disposal of such material contaminated by radioactive material

should be reviewed and modified where appropriate (Perma-Fix Environmental Services Inc.,

2015).

The New York State Department of Health (2014) review concluded that any risk assessment

should be supported with scientific information. It was considered that the currently

available scientific information on the hazards and risks associated with hydraulic fracturing

is insufficient to draw reliable conclusions. The authors concluded that until scientific

information allows accurate determination of the hazards and associated risks to the general

public, or the hazards and risks can adequately be managed, hydraulic fracturing should not

go ahead in New York State.

Air impacts

The studies examined in the review provided evidence of uncontrolled methane leaks, VOC

emissions and particulate matter being emitted from shale gas sites and the associated

infrastructure. The review cited recent studies in West Virginia (McCrawley, 2012, 2013; West

Virginia Department of Environmental Protection, 2013) in which traffic movements

associated with shale gas operations were cited as the likely source of high dust and benzene

levels. These increased levels were found to be present as far as 190 m from drilling sites.

Elevated benzene concentrations were also considered to have potentially originated from

the drilling and fracturing process. These emissions could contribute to health problems, e.g.,

respiratory problems, such as have been reported in areas of shale gas operations (e.g.,

Shonkoff et al., 2014).

Water quality

The studies reviewed by the New York State report have already been discussed in the water

contamination section (i.e., Warner et al., 2012; Warner et al., 2013; Darrah et al., 2014;

Vengosh et al., 2014). The review, however, fails to address the evidence for contamination

117

of shallow groundwater by natural methane migration over geological timescales as discussed

in the second briefing paper (i.e., Darrah et al., 2014).

Socioeconomic impacts

The New York State Department of Health review (2014) highlighted the risks associated with

rapid, concentrated expansion of extraction industries, such as precious minerals and

hydrocarbons, which result in the quality-of-life problems (e.g., noise and light pollution, and

odours) for the local population. Other potential issues highlighted in the review were

increased strain on the transport and health service infrastructure around areas of shale gas

operations. In addition, more rural areas will likely be unprepared for the increases in traffic

and population that are associated with shale gas operations. Such concerns have been raised

in previous studies (Texas Department of State Health Services, 2010a; Witter et al., 2010;

Stedman et al., 2012; West Virginia Department of Environmental Protection, 2013).

Health outcomes near sites

The health concerns raised by the present review have been alluded to previously in this

briefing document, specifically, skin rashes, nausea / vomiting, abdominal pain, breathing

difficulties / coughing, nosebleeds, anxiety / stress, headaches, dizziness, eye irritation, and

throat irritation (Bamberger and Oswald, 2012; Steinzor et al., 2012; Finkel and Hays, 2013).

The review highlighted a study by the National Institute for Occupational Safety and Health

(NIOSH), published by US Department of Labor, Occupational Safety and Health

Administration, in which sub-standard working practices and operational controls at well pads

contributed to levels of exposure to silica dust above those deemed safe by NIOSH (United

States Department of Labor Occupational Safety & Health Administration, 2012). The study

analysed 116 air samples from across 11 shale gas sites in Arkansas, Colorado, North Dakota,

Pennsylvania and Texas. 47% of the samples displayed silica exposure levels higher than those

determined safe for an 8-hour shift (0.1 mg/m3). If levels rise above 0.1 mg/m3, operators are

required to take action to reduce the level of exposure (United States Department of Labor

Occupational Safety & Health Administration, 2012). The study found that 9% of all samples

displayed silica exposure levels more than 10 times greater than the safe limit, with one

sample being 25 times greater. Samples that displayed elevated exposure levels came from

dust generation points, i.e., sand movers and blenders, and areas downwind of these points.

Some samples taken from upwind areas away from the dust generation points also displayed

elevated exposure levels; these were considered to be the result of truck movements. Vehicle

cabs without air conditioning and filtration displayed elevated exposure levels while those

with air conditioning and filtration did not.

118

High volume hydraulic fracturing health outcome studies

The New York State review found that the available information on the health impacts of

shale gas operations is limited and mainly exploratory. It was found that most studies fail

directly to demonstrate that exposure to contaminated substances, whether air, water or soil,

results in health effects. Furthermore, studies did not quantify the amount of exposure to

contaminated substances or demonstrate any direct causality between exposure and health

effects. Epidemiological studies were not found to be designed such as sufficiently to address

the association between shale gas operations and health effects whilst adequately controlling

bias and cofounding.

The review addresses a number of different areas of potential health impacts, these will be

summarised below.

Birth outcomes

The review cites studies by Hill (2013), which has been discussed at the start of the health

issues section of this document, and McKenzie et al. (2014) (which builds upon the (McKenzie

et al. (2012) study discussed previously in this briefing document). It should also be noted

that both of these studies did not quantify exposure to contaminated substances; instead

they used distance from the well pad as a proxy for exposure (New York State Department of

Health, 2014). The New York review points out that this is a reasonable approach for an initial

investigation. Further research should aim to quantify the exposure. The review also

highlights the fact that, by using distance from well pads as a proxy for exposure, the birth

outcome studies cannot identify specific risk factors and whether these risk factors were

related to shale gas operations (New York State Department of Health, 2014). For instance,

exposure to pollutants originating from everyday road traffic movements could not be

excluded from the studies.

Case series and symptom reports

Two of the studies cited by the New York State review (Bamberger and Oswald, 2012; Steinzor

et al., 2012) have previously been discussed. The review also cites a study by Rabinowitz et

al. (2015) in which the authors surveyed 492 people in 180 randomly selected homes located

across a range of distances from active drilling pads. The water supply for the properties

came from groundwater-fed wells. The proximity of each property to the drilling sites was

compared to the prevalence and frequency of medical symptoms, specifically, dermal,

respiratory, gastrointestinal, cardiovascular and neurological symptoms. The study found

that the number of reported symptoms per person increased with proximity to drilling sites.

Residents living <1 km from drilling sites reported an average of 3.27 symptoms per person

while those living >2 km from sites reported an average of 1.6 symptoms per person

(Rabinowitz et al., 2015). However, only the occurrence of upper respiratory symptoms

119

displayed a correlation with distance from wells (Rabinowitz et al., 2015); 39% of people living

in households located <1 km from wells reported symptoms while 31% of people living 1-2 km

from sites and 18% of those living >2 km from sites reported symptoms (Rabinowitz et al.,

2015).

Rabinowitz et al. (2015) and the New York State review (2014) point out that the study, and

those mentioned above, should be considered as hypothesis-generating, i.e., they suggest

possible relationships but more in-depth epidemiological studies with more rigorous controls

on factors such as bias, cofounding, temporality and chance findings. Further work would be

required in order to draw conclusions regarding the disease incidence and any casual

relationships between the proximity to shale gas operations and adverse health impacts. It is

also highlighted by the New York State review that, although the symptoms reported in these

studies are commonly reported in the wider general population, their apparently increased

incidence with proximity to shale gas sites means that the possibility of the observed

relationship to be due to shale gas activities cannot be ruled out. As pointed out above, the

current body of evidence prevents any firm conclusions being drawn on the issue.

Local community impacts

According to the New York State review (2014), there is a general consensus in the public

health community that certain social factors, known as social determinants of health, such as

income, education, housing and access to health care all influence health status. As

mentioned in the in the review by Werner et al. (2015), the rapid expansion of extraction

industries, such as shale gas, puts strain on not only the local health and transport

infrastructure due to the influx of new workers, but can also impact the quality of life of the

original local residents and can result in social problems.

One local community concern highlighted in the review is the number of trucks that will be

required to transport the water and other materials to the drilling sites. The review, citing

information from NTC Consultants (2011) suggest that approximately 1500 to 2000 truck trips

will be required over the lifetime of a single well – that is an average of 300 to 400 per year

over an assumed 20 year lifetime, with most movements taking place in the first year. Note

that this includes the transport of things like the drilling rig and chemical components of the

fracturing fluid to the site. The number of truck trips can be reduced considerably, however,

by increasing the well density per site and by piping the water into the site (NTC Consultants,

2011). In turn, this will decrease the impact associated with increased traffic, i.e., diesel

engine emissions, noise pollution and light pollution.

Increased road accidents are suggested to be a consequence of the increase in truck traffic

associated with the development of shale gas operations. The New York State review (2014)

cites a study by Graham et al. (2015), who found that the incidence of automobile and truck

120

accidents in 2010-2012 increased between 15-65% in counties with shale gas operations as

opposed to those without.

Cancer incidence

The review cites a study by Fryzek et al. (2013) which investigated the potential for association

between shale gas operations and childhood cancers in Pennsylvania. There was found to be

no increase in cancer after drilling had taken place when compared to the number of cancer

cases pre-drilling. The New York State review (2014) highlights limitations of the studies, i.e.,

rarity of childhood cancers and, more significantly, the lack of adequate lag time between the

onset of drilling and the emergence of any cancer cases. It should be noted that, even though

no increase was found, it may be too early for any increase to have taken place (New York

State Department of Health, 2014); therefore, future studies are required.

Shale gas Environmental Studies

Air quality impacts

The New York State review (2014) also cites studies conducted in Colorado (Colorado

Department of Public Health and Environment, 2010) and Texas (Texas Department of State

Health Services, 2010a; Bunch et al., 2014), the latter two of which have been discussed

previously. Additionally, the New York State review discusses the work by McKenzie et al.

(2012) on VOC exposure to people living at different distances from well sites. This study is

also discussed in the air contamination section of this document.

The New York State review (2014) also cites a study by Macey et al. (2014) measured air

samples from Arkansas, Colorado, Ohio, Pennsylvania and Wyoming. These samples were

taken from locations strategically identified through observations of industrial processes and

air impacts over the course of local residents’ daily routines. A total of 75 VOCs were tested

for. Of these, eight exceeded the federal guidelines for health based risk levels under several

operational circumstances (Macey et al., 2014). For example, in Wyoming, elevated levels of

hydrogen sulphide were found along the production chain, i.e., pump jacks, wastewater

discharge impounds and discharge canals. Of the eight compounds that exceeded federal

levels, benzene, formaldehyde and hydrogen sulphide were those that most commonly

exceed safe limits (Macey et al., 2014). The New York State review (2014) highlights the

shortcomings associated with the study, for instance, it was not clear whether the authors

had employed suitable risk-based comparison values. As an example, the review says that

the use of comparison values for long term cancer risk levels may have substantially

overstated the risk of cancer development associated with short term exposure levels that

were measured by Macey et al. (2014). There was also a lack of baseline measurements, e.g.,

upwind measurements and wind direction measurements (New York State Department of

Health, 2014). In addition, the review points out that, in some urban and industrial areas,

121

levels of benzene and formaldehyde commonly exceed the comparison values used by Macey

et al. (2014).

The Pennsylvania Department of Environmental Protection (PA DEP) carried out air sampling

during the hours at which most complaints related to shale gas activities were made, i.e., early

morning and late evening. The New York State report concentrates on results from north-

eastern and northcentral areas of Pennsylvania as these are the locations of natural gas

operations. The results found no immediate health risk to the general public (New York State

Department of Health, 2014). Some compounds, e.g., methyl mercaptan (a naturally

occurring compound found in some shales), were found in sufficient concentrations as to

result in odours (New York State Department of Health, 2014). The PA DEP indicate that

prolonged exposure odours, such as that produced by methyl mercaptan (the malodorous

component added to natural gas to make it easy for people to detect), can result in health

related impacts such as headaches and nausea. The New York State review did not give any

further details as what constitutes long term exposure.

The same PA DEP study measured carbon monoxide, nitrogen dioxide, sulphur dioxide and

ozone. Concentrations of these compounds did not exceed the National Ambient Air Quality

Standards in north-eastern Pennsylvania (Pennsylvania Department of Environmental

Protection, 2012). Benzene was also measured. Only one sample site indicated hazardous

levels (400 parts per billion) (Pennsylvania Department of Environmental Protection, 2012).

However, the monitoring device was set close to a car park, therefore the elevated

concentration can likely be attributed to automobile emissions (New York State Department

of Health, 2014). Another sampling site further away from the car park detected no elevated

benzene levels (Pennsylvania Department of Environmental Protection, 2012). The PA DEP

study concluded that benzene should not be considered a pollutant of concern near to

Marcellus Shale operations in Pennsylvania (Pennsylvania Department of Environmental

Protection, 2012).

The New York State review briefly cites evidence relating to increased benzene and alkane

levels measured at the National Oceanic and Atmospheric Administration’s Boulder

Atmospheric Observatory, Colorado, when the wind blew from the Denver-Julesburg Basin,

which is an area of oil and gas extraction (Pétron et al., 2012). This suggests that emissions

are being produced by the operations in the area. A study by Kemball-Cook et al. (2010)

documented elevated greenhouse gas and ozone emission levels in the area of the

Haynesville Shale, northeast Texas / northwest Louisiana. These elevated levels are

considered to be a result of oil and gas operations in the area through various methods

including hydraulic fracturing (Kemball-Cook et al., 2010).

Radon gas was also cited as a potential indoor contaminant. The New York review (2014) says

that a screening analysis by the Department of Health found that there is the potential for

radon from shale gas operations to contribute a small fraction of overall indoor radon

exposure although the review fails to quantify the fraction. Any potential contribution will

122

vary spatially as it has been demonstrated that there exists a large uncertainty regarding the

radon content of shales as there has been little in the way of radon monitoring (New York

State Department of Health, 2014).

Water Quality impacts

A large amount of the evidence cited by the New York State review has been discussed in

previously (e.g. Osborn et al., 2011; Darrah et al., 2014; Vengosh et al., 2014).

A study heavily cited by the New York State review is that written by Kassotis et al. (2014)

which has previously been discussed in the context of the PHE report (Kibble et al., 2013). In

addition to the shortcomings highlighted by Kibble et al. (2013), the New York State review

points out a number of additional shortcoming related to the study that limits the strength of

the conclusions. For instance, the source of the chemicals was not determined, therefore,

there is no guarantee that the elevated levels were a result of shale gas activities (New York

State Department of Health, 2014). In addition, the samples taken from drilling-dense and

reference areas were not correlated with each other for potentially influential factors, i.e.,

the type of water well (drinking or monitoring), depth of the well, stream ecology and

adjacent land use. This introduces inconsistencies in the data and fails to establish firm

controls on the baseline concentrations of the chemicals (New York State Department of

Health, 2014). The incidents that were associated with the sampling locations took place

months / years before the samples were taken and, in addition, no details of the bulk chemical

additives or the specific nature of the contamination event, or what remedial action was

taken, was provided (New York State Department of Health, 2014). The New York State

review says that, because of these shortcomings, the proximity of sample locations alone

cannot be used to indicate whether contamination incidents associated with shale gas

operations result in increased levels of endocrine disrupting chemicals; however, neither can

the possibility that be ruled out (New York State Department of Health, 2014). In agreement

with the PHE study, the New York State review points out that the implications for human

health raised by increased levels of EDCs in groundwater are limited based on the methods

used as the relevance of cell culture arrays to actual human exposure and human

physiological responses are unknown.

Induced earthquakes

The New York State review (2014) cites a study conducted in Oklahoma in which a swarm of

earthquakes, some of which were felt at the ground surface, were attributed to hydraulic

fracturing (Holland, 2014). In 2014, the Ohio Department of Natural Resources modified their

drilling permit conditions following an earthquake swarm in Poland Township where,

between March 4-12 2014, 77 earthquakes at magnitude 1-3 were detected (Skoumal et al.,

2015). These events were considered to be a result of the hydraulic fracturing process with

123

the largest events being a result of slip induced along a pre-existing fault / fracture zone

optimally orientated in the regional stress field (Skoumal et al., 2015). The seismic events at

Preese Hall are cited as evidence of hydraulic fracturing related seismic activity.

Conclusions from literature

The New York State review concluded, with regard to the reviewed health and environmental

literature, that current health and environmental studies are exploratory in nature and is best

viewed as hypothesis generating (New York State Department of Health, 2014). The authors

highlight the fact that many of the symptoms reported in the current health studies, i.e., skin

rashes, nausea / vomiting, abdominal pain, breathing difficulties / coughing, nosebleeds,

anxiety / stress, headaches, dizziness, eye irritation, and throat irritation are all acute or self-

limiting (New York State Department of Health, 2014). In addition, there are flaws with the

methods employed in studies examining birth weights and congenital defects, thus, limiting

the usefulness of the studies (New York State Department of Health, 2014). With regard to

the environmental literature, there are also problems with studies examining methane in

groundwater and fugitive emissions owing to the lack of background / baseline

measurements (New York State Department of Health, 2014). However, with regard to

factors such as earthquakes, there is no doubt that hydraulic fracturing can result in small

seismic events if the appropriate preliminary studies are not carried out (New York State

Department of Health, 2014).

Although the current studies that suggest some association between shale gas operations and

health impacts are inconclusive on account of the shortcomings discussed above, it is

important to consider that there may be significant health impacts that have not yet become

apparent. The New York review (2014) supports this by saying that adverse public health

impacts are largely unknown and that the current literature raises substantial questions about

whether enough is known about the public health risks as to allow them to be sufficiently

managed and mitigated.

Health Impact assessment

In addition to the published literature cited in the review, a number of health impact

assessments (HIA) made by various bodies in the US and the EU are discussed. Specifically,

the review cites HIA by the European Commission (Broomfield, 2012), Maryland Marcellus

Safe Drilling Initiative (University of Maryland Institute for Applied Environmental Health

School of Public Health, 2014), the University of Michigan (University of Michigan Graham

Sustainability Institute, 2013), the Research Triangle Environmental Health Cooperative

(Research Triangle Environmental Health Collaborative, 2013), the Nova Scotia Panel on

Hydraulic Fracturing (Wheeler et al., 2014), the National Institute of Environmental Health

Services (Penning et al., 2014) and the Institute of Medicine (Institute of Medicine of the

124

National Academies Board on Population Health and Public Health Practice Roundtable on

Environmental Health Sciences Research and Medicine (IOM), 2014). HIAs are structured

assessments that aim to assess health impacts and consequences of a policy, programme or

project. They are used by policy makers to inform future decisions and policy changes (Lock,

2000). HIAs are usually based on quantitative judgments when the issue in question is large-

scale, i.e., shale gas operations (New York State Department of Health, 2014).

Overall, the New York State review found that the concerns raised in these reports were

qualitatively similar to those raised in the academic literature. The review highlighted specific

public health risks emphasised in the HIAs. For instance, the European Commission

(Broomfield, 2012) considered that the cumulative risks associated with groundwater and

surface water contamination, depletion of groundwater, emissions and pollutants emitted to

the atmosphere, increased noise and increased traffic in the EU would all be “high”. The HIA

produced by the University of Michigan (University of Michigan Graham Sustainability

Institute, 2013) identified a range of health impacts including silica exposure, intentional

chemicals and chemicals produced during the hydraulic fracturing process, transportation, air

and water quality, ecological impacts (including the impact on recreational opportunities and

cultural / spiritual practices), and public perceptions resulting in health effects such as stress,

anxiety, depression etc. The North Carolina report (Research Triangle Environmental Health

Collaborative, 2013) highlighted the need for proper baselines to be established for water

quality, air quality and health statistics. Additionally, comprehensive water and wastewater

management plans are needed as well as an increasing level of regulation enforcement and

promotion of use of best practices (Research Triangle Environmental Health Collaborative,

2013). The National Institute of Environmental Health Services and the Institute of Medicine

(2014) reports both raised concerns about the potential for water and air pollution together

with the social disruption associated with the rapid expansion of shale gas operations.

Meetings with other States and consultation from medical professionals

The New York State review details the outcomes of meetings between representatives of the

New York Department of Health and representatives from agencies in other US states where

different approaches and risks where discussed. The States consulted were California

(Department of Public Health and the Department of Conservation), Texas (Department of

State Health Services, Railroad Commission and the Commission on Environmental Quality)

and Illinois (Department of Public Health and the Department of Natural Resources). In

addition, the Department of Health also consulted with external public health experts who

were asked to respond to three questions and provide any further comment. The questions

were (New York State Department of Health, 2014);

1. Are there additional potential public health impacts of high volume hydraulic

fracturing outside of those discussed in the draft supplemental generic environmental

impact statement (SGEIS)?

125

2. Are additional mitigation methods beyond those identified in the SGEIS needed to

address the potential health impacts of high volume hydraulic fracturing? If so, what

prevention or mitigation measures are recommended?

3. Are existing and proposed environmental and health monitoring surveillance systems

adequate to establish baseline health indicators and to measure potential health

impacts? If not, what additional monitoring is recommended?

The review contains letters addressing these questions from representatives from the

Colorado School of Public Health, The George Washington University and The University of

California, Los Angeles.

Common themes that run through the comments made by the state representatives and

external public health experts are air quality impacts, truck traffic impacts, noise pollution,

wastewater management, social disruption associated with rapidly-escalating

industrialization in communities, and cumulative effect of shale gas operations on stress (New

York State Department of Health, 2014). The data gaps in the currently available literature

pertaining to the indirect impact of shale gas operations on human health was highlighted by

public health experts (New York State Department of Health, 2014). These impacts are

considered to affect quality of life and stress due to factors including off-site nuisance odours

and visual pollution, e.g., light pollution (New York State Department of Health, 2014). The

lack of information regarding the impact of shale gas operations on surface waters and

wetlands via impacts on fish resources, other health related activities, e.g., swimming, and

flood control, were also raised (New York State Department of Health, 2014).

The HIAs also recognise the significant gaps in the public health knowledge relating to the

impact of high volume hydraulic fracturing on human health (New York State Department of

Health, 2014). The uncertainty of the effectiveness of some mitigation measures is also

pointed out. The need for a consistently evolving, robust regulatory system is also highlighted

together with the need for community engagement and promotion and enforcement of best

practices (New York State Department of Health, 2014).

Medact 2015 Report

A report published in 2015 by Medact (a British organisation made up of health care

professionals) addressed the health impacts and opportunity costs of hydraulic fracturing

(McCoy and Saunders, 2015). The report was produced in order to provide a broader

assessment of the potential impacts of hydraulic fracturing on public health and the quality

of the current regulatory system than have been presented in previous reports and reviews

(Broomfield, 2012; Kibble et al., 2013; Adgate et al., 2014; Cherry et al., 2014; Shonkoff et al.,

2014; University of Maryland Institute for Applied Environmental Health School of Public

Health, 2014; Wheeler et al., 2014; Werner et al., 2015). The authors of the study also

126

requested short papers from experts on specific subject area in addition to conducting

interviews with other academics and experts.

The report highlighted the potential sources of pollution produced during shale gas

operations, many of which have already been discussed previously. Specifically, the report

cites gas leaks that can occur throughout the extraction and supply chain, emissions from

diesel engines, silica being released into the atmosphere during vehicle movements, general

traffic movements, noise pollution, light pollution and odours (McCoy and Saunders, 2015).

Aside from the gas leaks, the remaining factors, which are associated with the rapid expansion

of industry, are considered to pose a significant risk to the health and wellbeing of local

communities (McCoy and Saunders, 2015). This will particularly be the case in situations

where the communities affected are rural or semi-rural (McCoy and Saunders, 2015).

With regard to traffic movements, McCoy and Saunders (2015) say that critical factors that

dictate the amount of traffic are the number of boreholes, whether water is piped or trucked

into the site, and the volume of wastewater that will need to be trucked away from the site.

There is also some variation in the estimations of the total truck movements over the lifetime

of a well. For instance, McCoy and Saunders (2015) cite estimates, given to the Environmental

Audit Committee’s Environmental Risks of Fracking Enquiry, from the Institute of Civil

Engineers (500-1,250 truck trips per well over a four week period, these trips just relate to

water transport) and the Royal Society for the Protection of Birds (4,300-6,600 truck trips per

well pad, this estimate included transport of equipment, fluid, sand and other materials

during the drilling, completion and hydraulic fracturing stages of the operation). The Medact

report considers that, if as many as 40 wells are located per well pad, then the total number

of truck trips could be as high as 34,000 over a time span of 2 years with the majority of the

movements taking place over the first 6 months. Two things should be noted regarding these

numbers. Firstly, 40 wells per well pad is a realistic upper estimate as this has been achieved

in the US and similar numbers may be seen in the UK if extraction goes ahead. A more realistic

estimate is 15 wells per well pad. Development of multi-well pads will be due to pressure on

operators to minimise the number of well pads. It should also be noted that the report makes

no mention of whether these truck movement numbers account for any water recycling that

would take place on-site. If it does not, it is likely that if recycling of flowback water took

place, the number of truck journeys would be reduced. Taking into account the initial

transport of steel pipe and drilling equipment, which need be moved only once, the number

of required truck journeys is not linearly proportional to the number of wells per pad. To put

these figures into perspective, UKOOG compares the number of truck journeys required per

year for shale gas production to the much larger number (370000 journeys per year) required

to move the milk produced by UK dairy herds.

McCoy and Saunders (2015) suggest that a further potential consequence of a large number

of truck movements is an increase in road traffic accidents. They cite the study by Graham

et al. (2015) discussed previously in the context of the New York State review (2014). Noise

127

pollution, light pollution and odours can all have an impact on the health, both physical and

mental, of the local population (McCoy and Saunders, 2015). The stress that all of the above

factors has on the local population can also not be discounted, as it has been shown that

stress can be a co-factor in the occurrence of other medical conditions (McCoy and Saunders,

2015, citing Gee and Payne-Sturges, 2004).

The gas itself will contain impurities and contaminants when it comes to the surface, these

are removed by either chemical scrubbing or flaring of the so-called “dirty gas” (McCoy and

Saunders, 2015). Oxides of nitrogen, hydrogen sulphide, formaldehyde, benzene, ethylene,

toluene, particulate matter and ground level ozone are noted as the main air-borne health

hazards (McCoy and Saunders, 2015). The degree of exposure to such pollutants will be

dependent upon a number of factors such as variations in geology and shale maturity, the

number of wells, the proximity to local population, topography, meteorological conditions,

what stage that operation is at, and operational practices (McCoy and Saunders, 2015).

McCoy and Saunders (2015) emphasise the need to consider the cumulative effect of multiple

wells when determining any potential health impacts. In terms of evidence for these impacts,

the report cites studies by McKenzie et al. (2012; 2014). These studies have been discussed

in previous sections of this document, although McCoy and Saunders (2015) fail to comment

on the shortcomings of either of these studies.

Surface and groundwater contamination by either gas, fracturing fluid or wastewater, is also

highlighted as a potential hazard. The report considers that between 10-90% of the fluid

pumped down the well during the fracturing stage will return to the surface. The Institute of

Civil Engineers considers 35% (approximately 7,500 to 18,750 m3 over the lifetime of the well)

to be a more typical value (McCoy and Saunders, 2015 citing written evidence to the

Environmental Audit Committee), however, as with all things shale gas-related, this will vary

from site to site. The authors propose that when shale formations are located deep

underground, the risk of groundwater contamination resulting from hydraulic fracturing

should be lower. However, they also speculate that if hydrofracturing takes place in a

geologically faulted area the risk of contamination remains. It should be noted that the

authors do not cite any evidence of this increased likelihood. Areas adjacent to those

fractured during injection and flowback are considered potentially to become contaminated

by methane or other gases (McCoy and Saunders, 2015). Fracturing fluid that does not return

to the surface is thought to be able to migrate into surrounding rock formations, it is also

presumed that aquifers can be contaminated with methane and other gases in areas adjacent

to hydraulic fracturing during the injection of fracturing fluid and during flowback (McCoy

and Saunders, 2015). The report provides no evidence for these claims and, in addition, no

mechanisms of gas / fluid migration are presented. McCoy and Saunders (2015) note that the

chance of contaminated water reaching members of the public will be lesser in the UK than

in the US due to the majority of drinking water being surface-derived and having undergone

treatment and quality control prior to entering into people’s homes. This is not the case in

128

the US, where many rural properties obtain their water from private, untreated drinking

wells.

Concerns about well integrity are also raised. A study by Kang et al. (2014) on methane

emissions from abandoned wells in Pennsylvania is cited as evidence of the problems with

long term well integrity. This Kang et al. study measured emissions from 19 abandoned oil

and gas wells (both plugged and unplugged), and so called “control areas” near the wells, in

late 2013 and early 2014. They found that all wells were producing more methane than the

control areas (0.27 kg per day per well (0.4 kg/m3 per day per well) of methane compared to

4.5x10-6 kg per day per well (1.0x10-5 kg/m3 per day per well) of methane). The emitted

methane was determined to be of thermogenic origin, i.e. from deep underground. Kang et

al. (2014) extrapolated their emissions measurements to account for the abandoned wells in

Pennsylvania and, in doing so, determined that approximately 4-7% of the total

anthropogenic methane emissions in Pennsylvania can be attributed to these wells. Well

integrity is also highlighted by McCoy and Saunders (2015) as being central to preventing

groundwater contamination. Citing studies discussed in previous (Davies et al., 2014;

Ingraffea et al., 2014; Jackson, 2014) McCoy and Saunders (2015) say that between 6-75% of

wells experience loss of barrier / well integrity or loss of zonal isolation with the level of failure

for unconventional wells being higher. McCoy and Saunders (2015) fail to provide potential

reasons why this is the case, however, causes have been discussed previously in this

document. In addition, the report cites the study by the Massachusetts Institute of

Technology (2011) discussed as part of the PHE review. However, the MIT report

acknowledges the fact that the data set is not comprehensive and should be used for

illustrative purposes to show the variety of potential incidents, the Medact report makes no

mention of this. The Medact report also says that potential interconnection between wells

and abandoned wells could provide a migration pathway for gases and fluids (McCoy and

Saunders, 2015, citing Kibble et al., 2013).

Social impacts noted by the report are related to the rapid expansion of extraction industry

in small rural or semi-rural communities. In addition, the ecology and aesthetics of the local

area may be altered. McCoy and Saunders (2015) acknowledge that, although growing, the

current evidence related to the effect of shale gas operations on ecosystems, agriculture and

animal husbandry is lacking (Bamberger and Oswald, 2012 is one of the few such studies).

The authors also highlight the need to consider wider social impacts associated with a rapidly

growing industry. For instance, they say that the influx of temporary workers (mostly young

men) has been shown to have a negative effect on the local community through increased

living costs, drug and alcohol use, mental illness and violence. However, the report does not

cite any evidence to support this claim. The county of Washington, Pennsylvania, has a decade

of experience of such population movements and impact son local communities as a result of

the development of their shale gas industry, and senior government officials have told the

task force that they did not encounter the problems speculated about in the Medact report.

129

The Medact report says that there are currently over 450 peer-reviewed publications in the

field of shale gas exploration and extraction. Despite this, there are still significant gaps in

the understanding of the level of health risk (McCoy and Saunders, 2015). The report cites

three main areas that require addressing before the full risk can be assessed. Firstly, the

toxicity of potential pollutants is yet to be quantified. For instance, McCoy and Saunders

(2015) say that benzene, which is known to be a potential pollutant produced during shale

gas operations, has no determined safe level of exposure. In addition, even though many of

the constituent chemicals used in the fracturing fluid have been allocated safety standards,

information about the effect of exposure to multiple chemicals is lacking. Note that the report

fails to cite any evidence for this; therefore, there may be no evidence of the impact of

exposure to multiple chemicals. Secondly, although hydraulic fracturing has been taking place

since the 1940s, the currently-used high volume hydraulic fracturing method is relatively new,

and as such, there are still few data and limited understanding about potential health impacts

(McCoy and Saunders, 2015). This is particularly the case regarding long term, robust

exposure studies (McCoy and Saunders, 2015). Because of this, the authors consider that the

accuracy with which the long term cumulative health effect can be predicted is low. Lastly,

the hazards and associated risks will vary from site-to-site based on geological, geographical,

social, demographic, social, agricultural and economic factors (McCoy and Saunders, 2015).

Not only will exposure vary from site-to-site, but also between individuals within the local

community. For instance, those members of the community that are more vulnerable,

whether through poor diet, deprivation or pre-existing health conditions, will likely be more

severely affected by shale gas operations (McCoy and Saunders, 2015). This adds a further

layer of complexity when trying to predict the health impact.

Other factors highlighted by the report that can affect the level of hazard and risk are the

number of wells per well pad, the density of well pads, the size and proximity of surrounding

communities, the location of any local aquifers supplying drinking water, operating

procedures of the operator, and the robustness and effectiveness of the regulatory system.

The review considers that if the density of boreholes is high in a small rural or semi-rural

location then the cumulative risk to the public health and wellbeing would be considerable.

However, the authors do not say whether they consider these risks to be unacceptable,

although one could consider that because McCoy and Saunders (2015) propose a

moratorium, this equates to these risks being unacceptable. UK-specific features that may

increase the hazards of shale gas exploration and extraction are a possible greater disposition

to seismic activity and presence of many geological faults in UK shales that might result in

damaged wells and pathways that might conduct environmental pollution. However, the

authors do not cite any firm evidence for this.

McCoy and Saunders (2015) acknowledge that it is impossible to mitigate completely all

hazards and associated risks from shale gas operations even if best practices are adopted and

suitably enforced due to the fact that accidents will happen. Some people will inevitably be

affected negatively by shale gas development whether it be economically (e.g., reduced

130

house prices) or socially (e.g., disruption to local infrastructure) (McCoy and Saunders, 2015).

Therefore, the question then becomes whether the hazards and risks can be sufficiently

mitigated as to make them acceptable, and whether the risks and benefits can be distributed

fairly. Overall, McCoy and Saunders (2015) consider that based on the gaps in current

understanding of the health risks associated with shale gas exploration and extraction,

together with uncertainties about the current regulatory system, that there should be a

minimal 5 year moratorium put in place on shale gas development in the UK until the health

and environmental impacts are understood better (McCoy and Saunders, 2015). But without

any exploratory activity from where will the required information come? Studies, presumably

to be carried out abroad, would have to account for all potential risks to health, i.e.,

cumulative and compound effects. Studies would have to be site specific with regard to

geological, economic, environmental and social factors, but would necessarily be non-UK

specific. Some extrapolation to full extraction would also have to be made in these studies.

Indeed, HIAs should contain this information as the HIA for small number of wells would be

inadequate if 40 wells were to be drilled at one site (McCoy and Saunders, 2015).

Additionally, McCoy and Saunders (2015) suggest that the information related to impacts on

the local community within HIAs should be presented in such a way that higher risk groups

can easily be distinguished from the majority of the population. Currently, there has been no

effort to examine the cumulative effect of shale gas extraction on an industrial scale in the UK

(McCoy and Saunders, 2015). Lastly, the Medact report recommends that such studies be

carried out by a body that is fully independent of the shale gas industry, although it is not

clear how this might be accomplished.

131

Preese Hall, Lancashire, Case Study

The Preese Hall drilling site, Lancashire, UK, is the most widely known UK site where

exploratory drilling has taken place for shale gas. The site was operated by Cuadrilla

Resources Ltd. The Preese Hall site has not moved past the exploration and appraisal stages

as defined by the United Kingdom Onshore Operators Group (2013a). The well has now been

shut-in and is being abandoned, with ongoing fugitive emissions monitoring taking place.

The drilling process, well integrity and hydraulic fracturing

Note: All information in this section was obtained from Cuadrilla Resources Ltd (2012). As

such, no references are included in this section.

Drilling began at the site in August of 2011 and reached a depth of 9097 ft. The schematic

diagram of the well assembly, together with the rock formations through which the drilling

took place, is shown in Figure 26. The well casing consisted of three main casing sections

(referred to as “holes” by Cuadrilla) and a conductor pipe. The conductor pipe was the first

casing section installed during the drilling process and extended down to the first solid layer

of bedrock (100 ft. at Preese Hall) upon which 20 inch casing cemented back to the surface.

The conductor pipe prevents near surface water, soil, gravel and sand from entering into the

borehole when the surface casing section is drilled. The surface casing section was first of the

three major sections drilled. Drilling continued down from the conductor pipe to a depth of

2021 ft., 800 ft. below the bottom of the Sherwood Sandstone aquifer (Figure 26). A 13-3/8

inch casing was then cemented back to the surface. The intermediate casing section was then

drilled to a depth of 4630 ft. and a 9-5/8 inch casing was inserted to a depth of 4603 ft. and

cemented back to the surface. The production casing section was the last section drilled.

Drilling went to a depth of 9097 ft., through the target formations of the Bowland and

Worston shale, and 5-1/2 inch casing was installed. This casing can withstand internal

pressures of 10,000 psi which is more than twice the pressure needed to induce hydraulic

fracturing, therefore, reducing the risk of casing failure during fracturing. At Preese Hall, the

lower part of this section was perforated in order to allow hydraulic fracturing to take place.

The intermediate casing section plays the largest role in determining well integrity as it passes

through the seal formation, in this case, the 400m thick Manchester marl (Figure 26). No

hydrocarbon deposits are found above this formation in the Bowland Basin, therefore, by

ensuring that the intermediate section is properly cemented in place, the risk of upwards flow

of drilling fluids and gases are reduced. Note, however, that the cement and casing need to

be installed correctly for this to occur.

Aside from these three main sections, a well cellar was also installed at the site. The well

cellar is a 3 m deep circular hole consisting of cement rings; these allow the blowout preventer

to be installed. The structure is leak-proof and allows the wellhead to be set into the ground,

thus minimising surface impact and aiding the process of site restoration.

132

Figure 26. Schematic diagram of the well drilled at the Preese Hall site together with the stratigraphic column of the local geology. Note that the diagram refers to the different casing sections as “hole” sections. (Cuadrilla Resources Ltd, 2012).

To minimise the risk of aquifer contamination, Cuadrilla focussed on ensuring that casing from

the intermediate casing section upwards was correctly cemented. This is because, in theory,

133

if the installation of cement and casing down to the base of the intermediate casing is of high

enough standard, the risk of ground water contamination is minimised as the cement should

form a seal with the Manchester marl.

Cuadrilla put in place a number of quality control measures to ensure that the cement and

casing were correctly installed. A suite of tests were first carried out on the cement

components (dry cement and the mixing water), this allowed the cement composition to be

determined. Before pumping began, a casing calliper log was run. This detects any areas of

the borehole that have become enlarged during drilling. The results of the log allow the

volume of cement to be determined. Cuadrilla pumped excess cement down the wellbore as

a safety measure to ensure that sufficient cement returned to the surface. The hole was

cleaned after drilling and prior to installation of the casing.

The cement was mixed and pumped into the casing using a purpose built, fully automated

machine. It constantly monitors the cement density, rate of mixing, injection pressures and

injection rates. To complement this, samples of the cement were taken every few minutes

and the cement mix density was physically measured using mud balance scales (industry

standard). During the pumping, Cuadrilla employed “basic quality control procedures” but

fail to state exactly what these procedures are and what they examine.

After each section was filled with cement, the operation was shut down for at least 24 hours

to allow the cement to dry and harden. The desired amount of hardening was determined

using the samples of cement collected during the pumping stage. Once the cement has set

to sufficient compressive strength, so as to allow drilling to re-commence without the cement

fracturing, the operation was restarted. When drilling was restarted, the entirety of the new

section was not drilled. Instead, a further 20 ft. of wellbore was drilled and a FIT carried out.

During drilling at Preese Hall, the average mud weight used during the drilling was 9.2 ppg to

10.5 ppg for the surface and intermediate sections. FIT tests were conducted at 12.5 ppg and

14.5 ppg respectively indicating that there was a 25% and 40% safety margin for the surface

and intermediate case respectively. Cuadrilla considered that these safety margins together

with the measures in place to ensure a high standard cement job reduce the risk of leakage

from the well to an acceptable level.

Cuadrilla are required to conduct a CBL through the surface casing if they observe that either

cement did not circulate back to the surface, if there were any problems with the mixing

and / or pumping of the cement or if the FIT indicates leakage at pressures lower than those

expected in the reservoir. A CBL was not carried out for the surface and intermediate casing

sections in Cuadrilla's first two onshore wells as none of these three events took place. They

have, however, now committed to conducting bond logs as often as possible on any future

wells drilled by Cuadrilla.

134

Wellbore integrity was found to be poor in multiple areas in the bottom 100 ft. of the

production casing. This was attributed to areas where the borehole was wider than expected

due to breakout. Before remedial action was taken, the areas of poor bonding were tested

to establish whether the casing was actually leaking. To do this, a plug was inserted below

the area of poor bonding and a number of perforations are made at the bottom and top of

the area. Another plug with a tube through the centre was inserted at the top of the area of

interest (Figure 27). Water was pumped, at pressure, through the small tube into the area of

low bonding. If the pressure remains constant then the cement is deemed acceptable and

hydraulic fracturing continues. If, on the other hand, pressure drops then it is implied that

the water is entering into the cement, therefore, the cement is in need of repair. The repair

is carried out by replacing the circulated water with cement (Figure 27). Once the

re-cementing has taken place and has had time to harden, the plugs are removed and the

well cleaned. This remedial action was carried out successfully at the Preese Hall site.

Multiple stage hydraulic fracturing from the bottom up took place at Preese Hall. A total of

six fracturing stages were carried out. This involved installing perforations in the section of

interest. Fracturing takes place and a casing plug is inserted which effectively seals off the

previously fractured section. The process was then repeated in the next area of interest.

135

Figure 27. Schematic diagram of the remedial action taken to repair the cement around the production casing at the Preese Hall site. The dark grey represents were bonded cement, the light grey represents medium bonded cement and the brown represents areas of poor bonding. (From Cuadrilla Resources Ltd, 2012)

Fluid usage and waste disposal

During drilling, Cuadrilla published details of the three chemicals they were given permission

to use in the drilling fluid (Department of Energy and Climate Change, 2014e; Cuadrilla

Resources Ltd., 2015)

136

Polyacrylamide – used to reduce friction. 0.043% used in the drilling fluid. This is a

non-hazardous, non-toxic substance commonly used in cosmetics as well as drinking

and wastewater plants.

Dilute hydrochloric acid –used to dilute the drilling fluid.

Glutaraldehyde biocide – used to cleanse and remove bacteria from the water.

The composition of the fracturing fluid can be seen in Appendix 2. Of the three chemicals,

only polyacrylamide was used as the water supplied by United Utilities and was found to be

sufficiently pure. In addition to polyacrylamide, a small amount of salt, which acts as a tracer,

was added the fluid (Cuadrilla Resources Ltd., 2015). Based on the chemical content of the

fluid, the EA classified the fracturing fluid as being non-hazardous (Department of Energy and

Climate Change, 2014e; Cuadrilla Resources Ltd., 2015).

The flowback water was tested at the ground surface by both the EA and Cuadrilla and was

found to contain small amounts of NORM. The water underwent treatment and disposed

according to EA regulations in the EA approved Davyhulme water treatment plants (Cuadrilla

Resources Ltd., 2015).

Seismic activity at Preese Hall

Cuadrilla used of ‘Buried Array’, provided by MicroSeismic, and ‘Tiltmeter Array’ provided by

Pinnacle Technologies will be installed in 104 specially prepared holes around the site

extending to depths ranging from only 12 to 90 metres. During the series of seismic events

that took place at the Preese Hall site in April of 2011, 55 earthquakes were recorded over

the course of two days (Figure 28). Prior to injection, no seismic activity was recorded at the

site. This, in conjunction with the close temporal correlation of the events with the injection

of the fluids (approximately 10 hours difference) led to the inference that the seismicity was

induced and not natural (Davies et al., 2013). The reason for this lag is considered to be either

that it represents the time needed for the fluid pressure to be transferred to the fault or that

the fault has some inherent storage and fluid flow capacity (Davies et al., 2013).

Earthquakes resulting from hydraulic fracturing

In April of 2011 two earthquakes of magnitude 1.5 and 2.3 were detected in the Blackpool

area by the BGS. DECC halted all fracturing at the site and a number of reports were

commissioned as to the cause of the earthquakes (de Pater and Baisch, 2011; Green et al.,

2012; Styles and Baptie, 2012). It was determined that the hydraulic fracturing at the Preese

Hall site was the cause of the earthquakes, specifically, movement of the fracturing fluids

along a critically stressed fault. It was, however, noted that similar magnitude seismic activity

resulting from coal mining had been documented in the area (Styles and Baptie, 2012). These

events occurred at shallower depths than those associated with hydraulic fracturing whilst

137

only causing minor damage. It is therefore reasonable to assume that any seismic events of

a similar magnitude originating from larger depth hydraulic fracturing would pose less risk.

An important note made by Styles and Baptie (2012) on the seismic events at the Preese Hall

site is that, although they agreed that the injection of fluid along the fractures resulted in the

earthquakes, the fact that the fault was critically stressed means that the tremors may have

taken place regardless sometime in the future.

Figure 28. Graph of the injected fluid volume and flowback fluid volume at the Preese Hall site in April 2011. Also displayed is the magnitude of the microseismic events associated with the injection. (Adapted from de Pater and Baisch, 2011 by Davies et al., 2013).

138

References Adam, R. and Kelsey, T. W. (2012). Pennsylvania dairy farms and Marcellus Shale, 2007–2010.

[Online] Available at :<http://pubs.cas.psu.edu/freepubs/pdfs/ee0020.pdf> [Accessed 20/03/2015]

Adams, M. B. (2011). Land application of hydrofracturing fluids damages a deciduous forest stand in West Virginia. Journal of Environmental Quality. 40(4), 1340-1344.

Adgate, J. L., Goldstein, B. D. and McKenzie, L. M. (2014). Potential public health hazards, exposures and health effects from unconventional natural gas development. Environmental Science & Technology. 48(15), 8307-8320.

Akcil, A. and Koldas, S. (2006). Acid Mine Drainage (AMD): causes, treatment and case studies. Journal of Cleaner Production. 14(12-13), 1139-1145.

Aksoy, N., Şimşek, C. and Gunduz, O. (2009). Groundwater contamination mechanism in a geothermal field: A case study of Balcova, Turkey. Journal of Contaminant Hydrology. 103(1-2), 13-28.

Alberta Energy Regulator. (2014). Flaring and Incineration. [online] Available at:<http://www.aer.ca/documents/enerfaqs/AER_EnerFAQs06_FlaringIncineration-Web.pdf> [Accessed 29/04/2015]

Allen, D. T., Torres, V. M., Thomas, J., Sullivan, D. W., Harrison, M., Hendler, A., Herndon, S. C., Kolb, C. E., Fraser, M. P., Hill, A. D., Lamb, B. K., Miskimins, J., Sawyer, R. F. and Seinfeld, J. H. (2013). Measurements of methane emissions at natural gas production sites in the United States. Proceedings of the National Academy of Sciences. 110(44), 17768-17773.

Almond, S., Clancy, S. A., Davies, R. J. and Worrall, F. (2014). The flux of radionuclides in flowback fluid from shale gas exploitation. Environmental Science and Pollution Research. 21(21), 12316-12324.

American Petroleum Institute. (2009). Environmental protection for onshore oil and gas production operations and leases; Recommended Practice 51R. Washington DC. [online] Available at:<https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/291485/LIT_7311_514be9.pdf> [Accessed 30/04/2015]

Arabasz, W. J., Nava, S. J., McCarter, M. K., Pankow, K. L., Pechmann, J. C., Ake, J. and McGarr, A. (2005). Coal-mining seismicity and ground-shaking hazard: A case study in the Trail Mountain area, Emery County, Utah. Bulletin of the Seismological Society of America. 95(1), 18-30.

Babiker, I. S., Mohamed, M. A. A., Terao, H., Kato, K. and Ohta, K. (2004). Assessment of groundwater contamination by nitrate leaching from intensive vegetable cultivation using geographical information system. Environment International. 29(8), 1009-1017.

Bamberger, M. and Oswald, R. E. (2012). Impacts of gas drilling on human and animal health. New solutions: a journal of environmental and occupational health policy. 22(1), 51-77.

Barbot, E., Vidic, N. S., Gregory, K. B. and Vidic, R. D. (2013). Spatial and temporal correlation of water quality parameters of produced waters from Devonian-age shale following hydraulic fracturing. Environmental Science & Technology. 47(6), 2562-2569.

Barker, J. F., Barbash, J. E. and Labonte, M. (1988). Groundwater contamination at a landfill sited on fractured carbonate and shale. Journal of Contaminant Hydrology. 3(1), 1-25.

BC Oil and Gas Commission. (2011). Oil and Gas Commission Fact Sheet; Overview: Flaring, Incinerating and Venting. [Online] Available at :< https://www.bcogc.ca/node/11464/download> [Accessed 29/04/2015]

Bhuiyan, M. A. H., Islam, M. A., Dampare, S. B., Parvez, L. and Suzuki, S. (2010). Evaluation of hazardous metal pollution in irrigation and drinking water systems in the vicinity of a coal mine area of north-western Bangladesh. Journal of Hazardous Materials. 179(1-3), 1065-1077.

139

Bott, R. D. (2007). Flaring: Questions + Answers. [Online] Available at :<http://siteresources.worldbank.org/EXTGGFR/Resources/578068-1258067586081/FlaringQA.pdf> [Accessed 29/04/2015]

Boyer, E. W., Swistock, B. R., Clark, J., Madden, M. and Rizzo, D. E. (2012). The Impact of Marcellus Gas Drilling on Rural Drinking Water Supplies. pp. 23

British Geological Survey. (2014). New Ollerton Earthquake Activity. [Online] Available at :<http://www.earthquakes.bgs.ac.uk/earthquakes/recent_events/ollerton_earthquakes.html> [Accessed 25/04/2015]

British Geological Survey. (2015a). Lower Greensand/Kimmeridge and Ampthill Clay separation map. [Online] Available at :<http://www.bgs.ac.uk/research/groundwater/shaleGas/aquifersAndShales/maps/separationMaps/lg-ka.html> [Accessed 01/05/2015]

British Geological Survey. (2015b). Triassic Sandstone/Bowland and Craven Groups separation map. [Online] Available at :<http://www.bgs.ac.uk/research/groundwater/shaleGas/aquifersAndShales/maps/separationMaps/ts-bcg.html> [Accessed 01/05/2015]

Broderick, J., Anderson, K., Wood, R., Gilbert, P., Sharmina, M., Footitt, A., Glynn, S. and Nicholls, F. (2011). Shale gas: an updated assessment of environmental and climate change impacts. A report commissions by the Co-operative and undertaken by researchers at the Tyndall Centre, University of Manchester. [Online] Available at :<http://www.tyndall.ac.uk/sites/default/files/coop_shale_gas_report_update_v3.10.pdf> [Accessed 03/05/2015]

Broni-Bediako, E. and Amorin, R. (2010). Effects of drilling fluid exposure to oil and gas workers presented with major areas of exposure and exposure indicators. Research Journal of Applied Sciences, Engineering and Technology. 2(8), 710-719.

Broomfield, M. (2012). Support to the Identification of Potential Risks for the Environment and Human Health Arising from Hydrocarbons Operations Involving Hydraulic Fracturing in Europe. European Commission DG Environment publication 07.0307/ENV.C.1/2011/604781/ENV.F1. [Online] Available at :<http://ec.europa.eu/environment/integration/energy/pdf/fracking%20study.pdf> [Accessed 19/04/2015]

Broomfield, M. and Donovan, B. (2012). Monitoring and control of fugitive methane from unconventional gas operations. pp. 93

Bunch, A. G., Perry, C. S., Abraham, L., Wikoff, D. S., Tachovsky, J. A., Hixon, J. G., Urban, J. D., Harris, M. A. and Haws, L. C. (2014). Evaluation of impact of shale gas operations in the Barnett Shale region on volatile organic compounds in air and potential human health risks. Science of the Total Environment. 468-469, 832-842.

Carbon Capture & Storage Association. (2015). What is CSS? [Online] Available at :<http://www.ccsassociation.org/what-is-ccs/> [Accessed 10/03/2015]

Carey, J. W., Svec, R., Grigg, R., Zhang, J. and Crow, W. (2010). Experimental investigation of wellbore integrity and CO2-brine flow along the casing-cement microannulus. International Journal of Greenhouse Gas Control. 4(2), 272-282.

Chapman, E. C., Capo, R. C., Stewart, B. W., Kirby, C. S., Hammack, R. W., Schroeder, K. T. and Edenborn, H. M. (2012). Geochemical and strontium isotope characterization of produced waters from Marcellus Shale natural gas extraction. Environmental Science & Technology. 46(6), 3545-3553.

Chartered Institute of Water and Environmental Management. (2014). Shale Gas and Water. [Online] Available at :<http://www.ciwem.org/media/1023226/Shale%20Gas%20and%20Water%20WEB.pdf> [Accessed 03/05/2015]

140

Cherry, J., Ben-Eli, M., Bharadwaj, L., Chalaturnyk, R., Dusseault, M., Goldstein, B., Lacoursière, J.-P., Matthews, R., Mayer, B., Molson, J., Munkittrick, K., Oreskes, N., Parker, B. and Young, P. (2014). Environmental Impacts of Shale Gas Extraction in Canada; The Expert Panel on Harnessing Science and Technology to Understand the Environmental Impacts of Shale Gas Extraction. [Online] Available at :<http://www.scienceadvice.ca/uploads/eng/assessments%20and%20publications%20and%20news%20releases/shale%20gas/shalegas_fullreporten.pdf> [Accessed 20/04/2015]

Choi, Y.-S., Young, D., Nešić, S. and Gray, L. G. S. (2013). Wellbore integrity and corrosion of carbon steel in CO2 geologic storage environments: A literature review. International Journal of Greenhouse Gas Control. 16, Supplement 1, S70-S77.

Chowdhury, S., Champagne, P. and McLellan, P. J. (2010a). Factorial analysis of trihalomethanes formation in drinking water. Water Environment Research. 82(6), 556-566.

Chowdhury, S., Champagne, P. and McLellan, P. J. (2010b). Investigating effects of bromide ions on trihalomethanes and developing model for predicting bromodichloromethane in drinking water. Water Research. 44(7), 2349-2359.

Classic, K. (2013). Radiation Basics. [Online] Available at :<http://hps.org/publicinformation/ate/faqs/radiation.html> [Accessed 10/06/2015]

Colborn, T., Kwiatkowski, C., Schultz, K. and Bachran, M. (2011). Natural Gas Operations from a Public Health Perspective. Human and Ecological Risk Assessment: An International Journal. 17(5), 1039-1056.

Colorado Department of Public Health and Environment. (2009). Garfield County emissions inventory. Air Pollution Control Division. pp. 34

Colorado Department of Public Health and Environment. (2010). Public health implications of ambient air exposures as measured in rural and urban oil & gas development areas — an analysis of 2008 air sampling data. pp. 29

Coons, T. and Walker, R. (2008). Community health risk analysis of oil and gas industry impacts in Garfield County. pp. 227

Coussens, C. and Martinez, R. M. (2013). Roundtable on Environmental Health Sciences, Research, and Medicine, The Board on Population Health and Public Health Practice, and the Institute of Medicine. Health impact assessment of shale gas extraction: workshop summary. Washington, DC. pp. 116

Crook, R. (2006). Cementing. In: Mitchell, R. F. (Eds). Petroleum Engineering Handbook, Volume II: Drilling Engineering. Society of Petroleum Engineers, 369-431.

Cuadrilla Resources Ltd. (2012). Wellbore integrity - Cuadrilla Land Based Wells. [Online] Available at :<http://www.cuadrillaresources.com/what-we-do/drilling/> [Accessed 01/02/2015]

Cuadrilla Resources Ltd. (2015). Water Disposal. [Online] Available at :<http://www.cuadrillaresources.com/protecting-our-environment/water/water-disposal/> [Accessed 18/06/2015]

Cuadrilla Resources Ltd. (2015). Fracturing Fluid. [Online] Available at :<http://www.cuadrillaresources.com/what-we-do/hydraulic-fracturing/fracturing-fluid/> [Accessed 27/01/2015]

Darrah, T. H., Vengosh, A., Jackson, R. B., Warner, N. R. and Poreda, R. J. (2014). Noble gases identify the mechanisms of fugitive gas contamination in drinking-water wells overlying the Marcellus and Barnett Shales. Proceedings of the National Academy of Sciences. 111(39), 14076-14081.

Davies, R., Foulger, G., Bindley, A. and Styles, P. (2013). Induced seismicity and hydraulic fracturing for the recovery of hydrocarbons. Marine and Petroleum Geology. 45, 171-185.

Davies, R. J. (2011). Methane contamination of drinking water caused by hydraulic fracturing remains unproven. Proceedings of the National Academy of Sciences. 108(43), E871.

141

Davies, R. J., Almond, S., Ward, R. S., Jackson, R. B., Adams, C., Worrall, F., Herringshaw, L. G., Gluyas, J. G. and Whitehead, M. A. (2014). Oil and gas wells and their integrity: Implications for shale and unconventional resource exploitation. Marine and Petroleum Geology. 56, 239-254.

Davies, R. J., Mathias, S. A., Moss, J., Hustoft, S. and Newport, L. (2012). Hydraulic fractures: How far can they go? Marine and Petroleum Geology. 37(1), 1-6.

Davis, S. D. and Frohlich, C. (1993). Did (or will) fluid injection cause earthquakes? - Criteria for a rational assessment. Seismological Research Letters. 64(3-4), 207-224.

de Pater, C. J. and Baisch, S. (2011). Geomechanical Study of Bowland Shale Seismicity: Synthesis Report. pp. 57

Department of Energy and Climate Change. (2013). Onshore oil and gas exploration in the UK: regulation and best practice. pp. 49

Department of Energy and Climate Change. (2014a). Fracking UK shale: local air quality. pp. 6 Department of Energy and Climate Change. (2014b). Fracking UK shale: regulation and monitoring.

pp. 6 Department of Energy and Climate Change. (2014c). Fracking UK shale: safety from design to

decommissioning. pp. 6 Department of Energy and Climate Change. (2014d). Fracking UK shale: understanding earthquake

risk. pp. 6 Department of Energy and Climate Change. (2014e). Fracking UK shale: water. pp. 8 DiGiulio, D. C., Wilkin, R. T., Miller, C. and Oberly, G. (2011). DRAFT: Investigation of Ground Water

Contamination near Pavillion, Wyoming. pp. 43 Dixon, D. W. (2001). Radon exposures from the use of natural gas in buildings. Radiation Protection

Dosimetry. 97(3), 259-264. Dogdu, M. S. and Bayari, C. S. (2005). Environmental impact of geothermal fluids on surface water,

groundwater and streambed sediments in the Akarcay Basin, Turkey. Environmental Geology. 47(3), 325-340.

Down, A., Armes, M. and Jackson, R. B. (2012). Shale gas extraction in North Carolina: research recommendations and public health implications. Environmental Health Perspectives. 121(10), A292.

EGS Solutions. (2015). Microseismic Monitoring 101. [Online] Available at :< https://www.esgsolutions.com/technical-resources/microseismic-knowledgebase/microseismic-monitoring-101> [Accessed 19/06/2015]

Engelder, T. (2012). Capillary tension and imbibition sequester frack fluid in Marcellus gas shale. Proceedings of the National Academy of Sciences. 109(52), E3625.

Environment Agency. (2010). Guidance for monitoring enclosed landfill gas flares. [Online] Available at :< https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/321610/LFTGN05.pdf> [Accessed 29/04/2015]

Environment Agency. (2011). Shale Gas. North West - Monitoring of Flowback water. [Online] Available at :<http://webarchive.nationalarchives.gov.uk/20140328084622/http://www.environment-agency.gov.uk/static/documents/Business/6th_Dec_-_Shale_gas_-_North_West_-_Monitoring_of_flowback_water_-_update_(3).pdf> [Accessed 18/06/2015]

Environment Agency. (2012a). Good practice for decommissioning redundant boreholes and wells. [Online] Available at :<http://webarchive.nationalarchives.gov.uk/20140328084622/http://cdn.environment-agency.gov.uk/LIT_6478_8cbe6f.pdf> [Accessed 22/04/2015]

Environment Agency. (2012b). Review of Assessment Procedures for Shale Gas Well Casing Installation. [Online] Available at :< https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/291485/LIT_7311_514be9.pdf> [Accessed 25/03/2015]

142

Environment Agency. (2013). An Environmental Risk Assessment for shale gas exploratory operations in England: Version 1. pp. 32

Environmental Audit Committee. (2015). Environmental risks of fracking. Eighth Report of Session 2014–15. pp. 42

Esswein, E. J., Breitenstein, M., Snawder, J., Kiefer, M. and Sieber, W. K. (2013). Occupational exposures to respirable crystalline silica during hydraulic fracturing. Journal of Occupational and Environmental Hygiene. 10(7), 347-356.

Evans, K. F., Zappone, A., Kraft, T., Deichmann, N. and Moia, F. (2012). A survey of the induced seismic responses to fluid injection in geothermal and CO2 reservoirs in Europe. Geothermics. 41, 30-54.

Finkel, M. L. and Hays, J. (2013). The implications of unconventional drilling for natural gas: A global public health concern. Public health. 127(10), 889-893.

Finkel, M. L., Hays, J. and Law, A. (2013a). Modern Natural Gas Development and Harm to Health: The Need for Proactive Public Health Policies. ISRN Public Health. 2013, 1-5.

Finkel, M. L., Selegean, J., Hays, J. and Kondamudi, N. (2013b). Marcellus Shale drilling's impact on the dairy industry in Pennsylvania: a descriptive report. New solutions: a journal of environmental and occupational health policy. 23(1), 189-201.

Fletcher, J. B. and McGarr, A. (2005). Moment tensor inversion of ground motion from mining-induced earthquakes, Trail Mountain, Utah. Bulletin of the Seismological Society of America. 95(1), 48-57.

Flewelling, S. A., Tymchak, M. P. and Warpinski, N. (2013). Hydraulic fracture height limits and fault interactions in tight oil and gas formations. Geophysical Research Letters. 40(14), 3602-3606.

Frac Focus. (2015a). Hydraulic Fracturing: The Process. [online] Available at:<https://fracfocus.org/hydraulic-fracturing-how-it-works/hydraulic-fracturing-process> [Accessed 02/04/2015]

Frac Focus. (2015b). Why are chemcials used? [online] Available at:<https://www.fracfocus.org/chemical-use/why-chemicals-are-used> [Accessed 06/01/2015]

Friends of the Earth. (2005). Gas Flaring in Nigeria. [online] Available at:<http://www.foe.co.uk/sites/default/files/downloads/gas_flaring_nigeria.pdf> [Accessed 18/06/2015]

Fryzek, J., Pastula, S., Jiang, X. and Garabrant, D. H. (2013). Childhood Cancer Incidence in Pennsylvania Counties in Relation to Living in Counties With Hydraulic Fracturing Sites. Journal of Occupational and Environmental Medicine. 55(7), 796-801.

Gee, G. C. and Payne-Sturges, D. C. (2004). Environmental Health Disparities: A Framework Integrating Psychosocial and Environmental Concepts. Environmental Health Perspectives. 112(17), 1645-1653.

Gibson, J. (2013). Summary Report of Review of Flare and Vent Gas Emissions Monitoring and Reporting Methods. A Report for National Measurment Office. pp. 26

Godwin, K. S., Hafner, S. D. and Buff, M. F. (2003). Long-term trends in sodium and chloride in the Mohawk River, New York: the effect of fifty years of road-salt application. Environmental Pollution. 124(2), 273-281.

Goldstein, B. D. and Malone, S. (2013). Obfuscation Does Not Provide Comfort: Response to the Article by Fryzek et al on Hydraulic Fracturing and Childhood Cancer. Journal of Occupational and Environmental Medicine. 55(11), 1376-1378.

Gordalla, B. C., Ewers, U. and Frimmel, F. H. (2013). Hydraulic fracturing: a toxicological threat for groundwater and drinking-water? Environmental Earth Sciences. 70(8), 3875-3893.

Graham, J., Irving, J., Tang, X., Sellers, S., Crisp, J., Horwitz, D., Muehlenbachs, L., Krupnick, A. and Carey, D. (2015). Increased traffic accident rates associated with shale gas drilling in Pennsylvania. Accident Analysis & Prevention. 74, 203-209.

143

Granberg, A. (2015). What Is Hydraulic Fracturing? [online] Available at:<http://www.propublica.org/special/hydraulic-fracturing-national> [Accessed 02/02/2015]

Green, C. A., Styles, P. and Baptie, B. (2012). Preese Hall shale gas fracturing: Review and recommendations for induced esismic mitigation. pp. 26

Halliburton. (2006a). Cementing: Bow-Spring Centralizers. [online] Available at:<http://www.halliburton.com/public/cem/contents/Data_Sheets/web/H/H05051-A4.pdf> [Accessed 27/03/2015]

Halliburton. (2006b). Cementing: Rigid Centralizers. [online] Available at:<http://www.halliburton.com/public/cem/contents/data_sheets/web/h/h05055.pdf> [Accessed 27/03/2015]

Halliburton. (2011). Microseismic Events. How Big Are These Microseismic Events? [online] Available at:<http://www.halliburton.com/public/pinnacle/contents/Data_Sheets/web/H08325.pdf> [Accessed 02/04/2015]

Hammack, R., Harbert, W., Sharma, S., Stewart, B., Capo, R., Wall, A., Wells, A., Diehl, R., Sams, J. and Veloski, G. (2014). An Evaluation of Fracture Growth and Gas/Fluid Migration as Horizontal Marcellus Shale Gas Wells are Hydraulically Fractured in Greene County, Pennsylvania; NETL-TRS-3-2014; EPAct Technical Report Series; U.S. Department of Energy, National Energy Technology Laboratory: Pittsburgh, PA. pp. 76

Healey, D. (2012). Hydraulic Fracturing or ‘Fracking’: A Short Summary of Current Knowledge and Potential Environmental Impacts. A Small Scale Study for the Environmental Protection Agency (Ireland) under the Science, Technology, Research & Innovation for the Environment (STRIVE) Programme 2007 – 2013. pp. 26

Health and Safety Executive. (2004). The selection and use of flammable gas detectors. pp. 16 Hill, E. L. (2013). Shale gas development and infant health: Evidence from Pennsylvania. [online]

Available at:<http://dyson.cornell.edu/research/researchpdf/wp/2012/Cornell-Dysonwp1212.pdf.> [Accessed 09/02/2015/

Hladik, M. L., Focazio, M. J. and Engle, M. (2014). Discharges of produced waters from oil and gas extraction via wastewater treatment plants are sources of disinfection by-products to receiving streams. Science of The Total Environment. 466-467(0), 1085-1093.

Holditch, S. A. (2006). Hydraulic fracturing. In: Clegg, D. C. (Eds). Petroleum Engineering Handbook, Volume IV: Production Operations Engineering. Society of Petroleum Engineers, 323-366.

Holland, A. A. (2014). Imaging Time Dependent Crustal Deformation Using GPS Geodesy and Induced Seismicity, Stress and Optimal Fault Orientations in the North American Mid-Continent. Graduate Thesis. University of Arizona. [online] Available at:<http://pqdtopen.proquest.com/doc/1614168780.html?FMT=AI> [Accessed 24/04/2015]

Horleston, A., Stork, A., Verdon, J., Baird, A., Wookey, J. and Kendall, M. (2013). Seismic Monitoring of Drilling Operations In Balcombe, West Sussex. [online] Available at:<http://www1.gly.bris.ac.uk/BUMPS/PDFS/BristolBalcombeReport2013.pdf> [Accessed 24/04/2015]

House of Commons. (2011). House of Commons Energy and Climate Change Committee. Shale Gas: Fifth Report of Session 2010–12; Seventh Special Report of Session 2010–12. [online] Available at:<https://www.gov.uk/government/uploads/system/uploads/attachment_data/file/79188/shalegas-govresp.pdf> [Accessed 03/05/2015]

House of Commons Energy and Climate Change Committee. (2011). Shale Gas Fifth Report of Session 2010–12. Volume I: Report, together with formal minutes, oral and written evidence pp. 65

Hubbert, M. K. and Willis, D. G. (1972). Mechanics of hydraulic fracturing(Eds). M 18: Underground Waste Management and Environmental Implications, 239-257.

Ingraffea, A. R. (2012). Fluid migration mechanisms due to faulty well design and/or construction: An overview and recent experiences in the Pennsylvania Marcellus play. pp. 10

144

Ingraffea, A. R., Wells, M. T., Santoro, R. L. and Shonkoff, S. B. C. (2014). Assessment and risk analysis of casing and cement impairment in oil and gas wells in Pennsylvania, 2000–2012. Proceedings of the National Academy of Sciences. 111(30), 10955-10960.

Institute of Medicine of the National Academies Board on Population Health and Public Health Practice Roundtable on Environmental Health Sciences Research and Medicine (IOM). (2014). Health Impact Assessment of Shale Gas Extraction: Workshop Summary. [online] Available at:<http://www.nap.edu/catalog/18376/health-impact-assessment-of-shale-gas-extraction-workshop-summary> [Accessed 19/04/2015]

IPIECA. (2013). Green Completions. [online] Available at:<http://www.ipieca.org/energyefficiency/solutions/78161> [Accessed 26/01/2015]

Jackson, R. B. (2014). The integrity of oil and gas wells. Proceedings of the National Academy of Sciences. 111(30), 10902-10903.

Jackson, R. B., Vengosh, A., Darrah, T. H., Warner, N. R., Down, A., Poreda, R. J., Osborn, S. G., Zhao, K. and Karr, J. D. (2013). Increased stray gas abundance in a subset of drinking water wells near Marcellus shale gas extraction. Proceedings of the National Academy of Sciences. 110(28), 11250-11255.

Johnson, D. B. and Hallberg, K. B. (2005). Acid mine drainage remediation options: a review. Science of The Total Environment. 338(1-2), 3-14.

Johnson, R. (1973). Assessment of Potential Radiological Health Effects from Radon in Natural Gas. Report for US Environmental Protection Agency. United States Environmental Protection Agency. [online] Available at:<http://rwma.com/Johnson%20et%20al.,%201973.pdf> [Accessed 02/05/2015]

Kaktins, N. M. (2011). Drilling the Marcellus shale for natural gas: environmental health issues for nursing. The Pennsylvania nurse. 66(1), 4.

Kang, M., Kanno, C. M., Reid, M. C., Zhang, X., Mauzerall, D. L., Celia, M. A., Chen, Y. and Onstott, T. C. (2014). Direct measurements of methane emissions from abandoned oil and gas wells in Pennsylvania. Proceedings of the National Academy of Sciences. 111(51), 18173-18177.

Kargbo, D. M., Wilhelm, R. G. and Campbell, D. J. (2010). Natural Gas Plays in the Marcellus Shale: Challenges and Potential Opportunities. Environmental Science & Technology. 44(15), 5679-5684.

Kassotis, C. D., Tillitt, D. E., Davis, J. W., Hormann, A. M. and Nagel, S. C. (2014). Estrogen and androgen receptor activities of hydraulic fracturing chemicals and surface and ground water in a drilling-dense region. Endocrinology. 155(3), 897-907.

Kemball-Cook, S., Bar-Ilan, A., Grant, J., Parker, L., Jung, J., Santamaria, W., Mathews, J. and Yarwood, G. (2010). Ozone Impacts of Natural Gas Development in the Haynesville Shale. Environmental Science & Technology. 44(24), 9357-9363.

Kibble, A., Cabianca, T., Daraktchieva, Z., Gooding, T., Smithard, J., Kowalczyk, G., McColl, N. P., Singh, M., Mitchem, L., Lamb, P., Vardoulakis, S. and Kamanyire, R. (2013). Public Health England: Review of the Potential Public Health Impacts of Exposures to Chemical and Radioactive Pollutants as a Result of the Shale Gas Extraction Process. pp. 52

King, G. E. (1996). An Introduction to the Basics of Well Completions, Stimulations and Workovers; Chapter 3: Cement. 2nd ed.

King, G. E. and King, D. E. (2013). Environmental risk arising from well-construction failure - Differences between barrier and well failure, and estimates of failure frequency across common well types, locations, and well age. SPE Production & Operations. 28(4), 323-344.

Kinik, K. and Wojtanowicz, A. K. (2011). Identifying environmental risk of sustained casing pressure. Society of Petroleum Engineers Conference Paper. SPE-143713-MS.

Kondash, A. J., Warner, N. R., Lahav, O. and Vengosh, A. (2014). Radium and barium removal through blending hydraulic fracturing fluids with acid mine drainage. Environmental Science & Technology. 48(2), 1334-1342.

145

La Plata County. (2002). La Plata County impact report. [online] Available at:<http://www.co.laplata.co.us/departments_elected_officials/planning/natural_resources_oil_gas/impact_report> [Accessed 20/03/2015]

Laney, A. S. and Weissman, D. N. (2012). The Classic Pneumoconioses: New Epidemiological and Laboratory Observations. Clinics in Chest Medicine. 33(4), 745-758.

Levallois, P., Thériault, M., Rouffignat, J., Tessier, S., Landry, R., Ayotte, P., Girard, M., Gingras, S., Gauvin, D. and Chiasson, C. (1998). Groundwater contamination by nitrates associated with intensive potato culture in Québec. Science of The Total Environment. 217(1-2), 91-101.

Lock, K. (2000). Health impact assessment. BMJ : British Medical Journal. 320(7246), 1395-1398. Logan, J., Heath, G., Macknick, J., Paranhos, E., Boyd, W. and Carlson, K. (2012). The Joint Institute for

Strategic Energy Analysis Report. Natural Gas and the Transformation of the U.S. Energy Sector: Electricity. pp. 240

Løseth, H., Wensaas, L., Arntsen, B., Hanken, N.-M., Basire, C. and Graue, K. (2011). 1000 m long gas blow-out pipes. Marine and Petroleum Geology. 28(5), 1047-1060.

Macey, G., Breech, R., Chernaik, M., Cox, C., Larson, D., Thomas, D. and Carpenter, D. (2014). Air concentrations of volatile compounds near oil and gas production: a community-based exploratory study. Environmental Health. 13(1), 82.

MacKay, D. J. C. and Stone, T. J. (2013). Potential Greenhouse Gas Emissions Associated with Shale Gas Extraction and Use. pp. 39

Massachusetts Institute of Technology. (2011). Study on the Future of Natural Gas. MIT Energy Initiative. [online] Available at:<http://mitei.mit.edu/publications/reports-studies/future-natural-gas> [Accessed 21/04/2015]

McCoy, D. and Saunders, P. (2015). Health and Fracking: The impact and opportunity costs. pp. 32 McCrawley, M. (2012). Air, Noise, and Light Monitoring Plan for Assessing Environmental Impacts of

Horizontal Gas Well Drilling Operations (ETD-10 Project), West Virginia University School of Public Health, Morgantown, WV. [online] Available at:<http://www.dep.wv.gov/oil-and-gas/Horizontal-Permits/legislativestudies/Documents/WVU%20Final%20Air%20Noise%20Light%20Protocol.pdf> [Accessed 10/04/2015]

McCrawley, M. (2013). Air, Noise, and Light Monitoring Results for Assessing Environmental Impacts of Horizontal Gas Well Drilling Operations. West Virginia University School of Public Health, Morgantown, WV. [online] Available at:<http://wvwri.org/wp-content/uploads/2013/10/a-n-l-final-report-for-web.pdf> [Accessed 10/04/2015]

McGarr, A. and Fletcher, J. B. (2005). Development of ground-motion prediction equations relevant to shallow mining-induced seismicity in the Trail Mountain Area, Emery County, Utah. Bulletin of the Seismological Society of America. 95(1), 31-47.

McKenzie, L. M., Guo, R., Witter, R. Z., Savitz, D. A., Newman, L. S. and Adgate, J. L. (2014). Birth outcomes and maternal residential proximity to natural gas development in rural Colorado. Environ Health Perspect. 122(4), 412-7.

McKenzie, L. M., Witter, R. Z., Newman, L. S. and Adgate, J. L. (2012). Human health risk assessment of air emissions from development of unconventional natural gas resources. Science of The Total Environment. 424, 79-87.

Microseismic. (2014). Trusted Data Acquisition Methods: Enhacning Reservoir Completions Techniques and Production. [online] Available at:<http://www.microseismic.com/brochures/MSI_Brochure_DataAcquisition.pdf> [Accessed 19/06/2015]

Miyazaki, B. (2009). Well integrity: An overlooked source of risk and liability for underground natural gas storage. Lessons learned from incidents in the USA. Geological Society, London, Special Publications. 313(1), 163-172.

146

Molofsky, L. J., Connor, J. A., Farhat, S. K., Wylie, A. S. and Wagner, T. (2011). Methane in Pennsylvania water wells unrelated to Marcellus shale fracturing. Oil and Gas Journal. 109(49), 54-67.

Molofsky, L. J., Connor, J. A., Wylie, A. S., Wagner, T. and Farhat, S. K. (2013). Evaluation of Methane Sources in Groundwater in Northeastern Pennsylvania. Groundwater. 51(3), 333-349.

Monteiro Santos, F. A., Mateus, A., Figueiras, J. and Gonçalves, M. A. (2006). Mapping groundwater contamination around a landfill facility using the VLF-EM method - A case study. Journal of Applied Geophysics. 60(2), 115-125.

Montgomery, C. (2013). Fracturing Fluids. In: Bunger, A. P., McLennan, J. & Jeffrey, R. (Eds). Effective and Sustainable Hydraulic Fracturing, 1-24.

Mor, S., Ravindra, K., Dahiya, R. P. and Chandra, A. (2006). Leachate Characterization and Assessment of Groundwater Pollution Near Municipal Solid Waste Landfill Site. Environmental Monitoring and Assessment. 118(1-3), 435-456.

Murat Özler, H. (2000). Hydrogeology and geochemistry in the Curuksu (Denizli) hydrothermal field, western Turkey. Environmental Geology. 39(10), 1169-1180.

Myers, T. (2012). Potential contaminant pathways from hydraulically fractured shale to aquifers. Groundwater. 50(6), 872-882.

National Commission on the BP Deepwater Horizon Oil Spill and Offshore Drilling. (2011). Macondo: The Gulf Oil Disaster; Chief Counsel’s Report. Chapter 4.6. [online] Available at:<http://www.eoearth.org/files/164401_164500/164441/c21462-214_ccr_chp_4-6_negative_pressure_test.pdf> [Accessed 18/06/2015]

National Energy Technology Laboratory. (2014). Development of Nanoparticle-Stabilized Foams To Improve Performance of Water-less Hydraulic Fracturing. [online] Available at:<http://www.netl.doe.gov/research/oil-and-gas/project-summaries/enhanced-oil-recovery/fe0013723-UTA> [Accessed 28/04/2015]

National Energy Technology Laboratory. (2015a). Development and Field Testing Novel Natural Gas Surface Process Equipment for Replacement of Water as Primary Hydraulic Fracturing Fluid. [online] Available at:<http://www.netl.doe.gov/research/oil-and-gas/project-summaries/unconventional-resources/fe0024314-swri> [Accessed 28/04/2015]

National Energy Technology Laboratory. (2015b). Methods to Enhance Wellbore Cement Integrity with Microbially-Induced Calcite Precipitation (MICP). [online] Available at:<http://www.netl.doe.gov/research/oil-and-gas/project-summaries/unconventional-resources/fe0024296-msu> [Accessed 28/04/2015]

National Energy Technology Laboratory. (2015c). “Nanite” for Better Well-Bore Integrity and Zonal Isolation. [online] Available at:<http://www.netl.doe.gov/research/oil-and-gas/project-summaries/natural-gas-resources/fe0014144-Oceanit> [Accessed 28/04/2015]

National Energy Technology Laboratory. (2015d). nXis Well Integrity Inspection in Unconventional Wells. [online] Available at:<http://www.netl.doe.gov/research/oil-and-gas/project-summaries/unconventional-resources/fe0024293-geglobal> [Accessed 28/04/2015]

National Energy Technology Laboratory. (2015e). Unconventional Resources. [online] Available at:<http://www.netl.doe.gov/research/oil-and-gas/project-summaries/unconventional-resources> [Accessed 28/04/2015]

National Institute of Environmental Health Sciences. (2015). Endocrine Disruptors. [online] Available at:<http://www.niehs.nih.gov/health/topics/agents/endocrine/> [Accessed 15/06/2015]

Natural Gas STAR Partners. (2006a). Optimize Glycol Circulation And Install Flash Tank Separators In Glycol Dehydrators pp. 10

Natural Gas STAR Partners. (2006b). Options For Reducing Methane Emissions From Pneumatic Devices In The Natural Gas Industry pp. 12

Natural Gas STAR Partners. (2006c). Replacing Glycol Dehydrators With Desiccant Dehydrators pp. 14 New York State Department of Environmental Conservation. (2011). Revised Draft SGEIS on the Oil,

Gas and Solution Mining Regulatory Program (September 2011). Well Permit Issuance for

147

Horizontal Drilling and High-Volume Hydraulic Fracturing in the Marcellus Shale and Other Low-Permeability Gas Reservoirs. pp. 1537

New York State Department of Health. (2014). A Public Health Review of High Volume Hydraulic Fracturing for Shale Gas Development. pp. 176

Neymeyer, A., Williams, R. T. and Younger, P. L. (2007). Migration of polluted mine water in a public supply aquifer. Quarterly Journal of Engineering Geology and Hydrogeology. 40(1), 75-84.

Nolen-Hoeksema, R. (2013). Elements of hydraulic fracturing. Oilfield Review. 25(2), 51-52. NTC Consultants. (2011). Impacts on community character of horizontal drilling and high volume

hydraulic fracturing in Marcellus shale and other low-permeability gas reservoirs. Final Report. Prepared for the New York State Energy Research and Development Authority. pp. 45

Oil and Gas UK. (2009). Guidance on suspension and abandonment of wells/ North Sea well abandonment study. OP006. Issue 3. London: Oil and Gas UK.

Oil and Gas UK. (2012). Well integrity guidelines. [online] Available at:<http://www.oilandgasuk.co.uk/cmsfiles/modules/publications/pdfs/OP069.pdf> [Accessed 10/06/2015]

Osborn, S. G., Vengosh, A., Warner, N. R. and Jackson, R. B. (2011). Methane contamination of drinking water accompanying gas-well drilling and hydraulic fracturing. Proceedings of the National Academy of Sciences. 108(20), 8172-8176.

OSPAR Commission. (2012). Discharges, spills and emissions from offshore oil and gas installations in 2010. pp. 48

Otton, J. K. (1992). Radon in soil gas and soil radioactivity in Prince George's County. [online] Available at:<http://pubs.usgs.gov/of/1992/0011/report.pdf> [Accessed 02/05/2015]

Ovuakporaye, S. I., Aloamaka, C. P., Ojieh, A. E., Ejebe, D. E. and Mordi, J. C. (2012). Effects of gas flaring on lung function among residents in gas flaring community in Delta State, Nigeria. Research Journal of Environmental and Earth Sciences. 4(5), 525-528.

Papoulias, D. M. and Velasco, A. L. (2013). Histopathological analysis of fish from Acorn Fork Creek, Kentucky exposed to hydraulic fracturing fluid releases. Southeastern Naturalist. 12(4).

Penning, T. M., Breysse, P. N., Gray, K., Howarth, M. and Yan, B. (2014). Environmental health research recommendations from the inter-environmental health sciences core center working group on unconventional natural gas drilling operations. Environmental Health Perspectives. 122(11), 1155.

Pennsylvania Department of Environmental Protection. (2012). DEP Conducting Long-Term Marcellus Shale Air Monitoring Study in Southwestern Pennsylvania. [online] Available at:<http://www.portal.state.pa.us/portal/server.pt/community/oil_and_gas_related_topics/20349/air/986695> [Accessed 14/04/2015]

Perma-Fix Environmental Services Inc. (2015). Technologically Enhanced Naturally Occurring Radioactive Materials (TENORM) Study Report: A report prepared for teh Pennsylvania Department of Environmental Protection. [online] Available at:<http://www.elibrary.dep.state.pa.us/dsweb/Get/Document-105822/PA-DEP-TENORM-Study_Report_Rev._0_01-15-2015.pdf> [Accessed 10/04/2015]

Perry, S. L. (2012). Environmental reviews and case studies: Addressing the societal costs of unconventional oil and gas exploration and production: A framework for evaluating short-term, future, and cumulative risks and uncertainties of hydrofracking. Environmental Practice. 14(4), 352-365.

Perry, S. L. (2013). Using ethnography to monitor the community health implications of onshore unconventional oil and gas developments: examples from Pennsylvania's Marcellus Shale. New solutions: a journal of environmental and occupational health policy. 23(1), 33-53.

Pétron, G., Frost, G., Miller, B. R., Hirsch, A. I., Montzka, S. A., Karion, A., Trainer, M., Sweeney, C., Andrews, A. E., Miller, L., Kofler, J., Bar-Ilan, A., Dlugokencky, E. J., Patrick, L., Moore, C. T., Ryerson, T. B., Siso, C., Kolodzey, W., Lang, P. M., Conway, T., Novelli, P., Masarie, K., Hall, B., Guenther, D., Kitzis, D., Miller, J., Welsh, D., Wolfe, D., Neff, W. and Tans, P. C. D. (2012).

148

Hydrocarbon emissions characterization in the Colorado Front Range: A pilot study. Journal of Geophysical Research: Atmospheres. 117(D4), n/a-n/a.

PetroWiki. (2014a). Cement bond logs. [online] Available at:<http://petrowiki.org/Cement_bond_logs#Noteworthy_papers_in_OnePetro> [Accessed 12/06/2015]

PetroWiki. (2014b). Cement composition and classification. [online] Available at:<http://petrowiki.org/Cement_composition_and_classification#API_classifications> [Accessed 25/03/2015]

PetroWiki. (2014c). Cementing hardware. [online] Available at:<http://petrowiki.org/Cementing_hardware> [Accessed 27/03/2015]

Petrowiki. (2015). Gas lift. [online] Available at:<http://petrowiki.org/Gas_lift> [Accessed 29/04/2015]

Rabinowitz, P. M., Slizovskiy, I. B., Lamers, V., Trufan, S. J., Holford, T. R., Dziura, J. D., Peduzzi, P. N., Kane, M. J., Reif, J. S. and Weiss, T. R. (2015). Proximity to natural gas wells and reported health status: Results of a household survey in Washington County, Pennsylvania. Environmental Health Perspectives. 123(1), 21-26.

Rassenfoss, S. (2011). From flowback to fracturing: water recycling grows in the Marcellus shale. Journal of Petroleum Technology. 63(7), 48-51.

ReFINE. (2014). How radioactive is fracking flowback water? DEI Briefing Note No. 905: Autumn 2014. [online] Available at:<https://www.dur.ac.uk/resources/refine/NORMRB.pdf> [Accessed 27/04/2015]

Research Partnership to Secure Energy for America. (2013). Development of Non-Contaminating Cryogenic Fracturing Technology for Shale and Tight Gas Reservoirs. [online] Available at:<http://www.rpsea.org/media/files/project/3d9af270/10122-20-Development_Non-Contaminating_Cryogenic_Fracturing_Technology_Shale_Tight_Gas_Reservoirs-06-17-13.pdf> [Accessed

Research Partnership to Secure Energy for America. (2014). Prevention and Remediation of Sustained Casing Pressure and other Isolation Breaches. [online] Available at:<http://www.rpsea.org/media/files/project/b0d66da0/11122-42-PFS-Prevention_Remediation_Sustained_Casing_Pressure_Isolation_Breaches-03-19-14.pdf> [Accessed

Research Triangle Environmental Health Collaborative. (2013). Shale Gas Extraction in North Carolina: Public Health Implications. [online] Available at:<http://environmentalhealthcollaborative.org/images/2012SummitWorkProduct.pdf> [Accessed 19/04/2015]

Resnikoff, M. (2012). Radon in Natural Gas from Marcellus Shale. [online] Available at:<http://rwma.com/Marcellus%20Radon-%20Ethics.pdf> [Accessed 02/05/2015]

Resnikoff, M., Alexandrova, E. and Travers, J. (2010). Radioactivity in Marcellus Shale Report prepared for Residents for the Preservation of Lowman and Chemung (RFPLC) [online] Available at:<http://energy.wilkes.edu/PDFFiles/Library/Marcellus%20Shale%20Radioactivity%20Report%205-18-2010.pdf> [Accessed

Rich, A., Grover, J. P. and Sattler, M. L. (2014). An exploratory study of air emissions associated with shale gas development and production in the Barnett Shale. Journal of the Air & Waste Management Association. 64(1), 61-72.

RKI Instruments. (2014). What is %LEL and %UEL. [online] Available at:<http://rkiinstruments.com/pages/faq/What_is_LEL_UEL.htm> [Accessed 18/03/2015]

Rowan, E. L. and Kraemer, T. F. (2012). Radon-222 Content of Natural Gas Samples from Upper and Middle Devonian Sandstone and Shale Reservoirs in Pennsylvania: Preliminary Data. US Geological Survey Open-file Report 2012–1159. [online] Available at:<http://pubs.usgs.gov/of/2012/1159/ofr2012-1159.pdf> [Accessed 02/05/2015]

149

Roy, A. A., Adams, P. J. and Robinson, A. L. (2014). Air pollutant emissions from the development, production, and processing of Marcellus Shale natural gas. Journal of the Air & Waste Management Association. 64(1), 19-37.

Saberi, P. (2013). Navigating medical issues in shale territory. New solutions: a journal of environmental and occupational health policy. 23(1), 209-221.

Saiers, J. E. and Barth, E. (2012). Comment on "Potential contaminant pathways from hydraulically fractured shale aquifers". Groundwater. 50(6), 826-828.

Searl, A. and Galea, K. S. (2011). Toxicological review of the possible effects associated with inhalation and dermal exposure to drilling fluid production stream. pp. 84

Shariq, L. (2013). Uncertainties associated with the reuse of treated hydraulic fracturing wastewater for crop irrigation. Environmental Science & Technology. 47(6), 2435-2436.

Sherar, B. W. A., Power, I. M., Keech, P. G., Mitlin, S., Southam, G. and Shoesmith, D. W. (2011). Characterizing the effect of carbon steel exposure in sulfide containing solutions to microbially induced corrosion. Corrosion Science. 53(3), 955-960.

Shonkoff, S. B. C., Hayes, J. and Finkel, M. L. (2014). Environmental public health dimensions of shale and tight gas development. Environmental Health Perspectives. 122(8), 787-795.

Singh, R. P. and Yadav, R. N. (1995). Prediction of subsidence due to coal mining in Raniganj coalfield, West Bengal, India. Engineering Geology. 39(1), 103-111.

Skoumal, R. J., Brudzinski, M. R. and Currie, B. S. (2015). Earthquakes induced by hydraulic fracturing in Poland Township, Ohio. Bulletin of the Seismological Society of America.

Sperger, C., Cook, K. and Klemow, K. (2012). Does Marcellus Shale pose a radioactivity risk? Institute for Energy and Environmental Research of Northeastern Pennsylvania Clearinghouse. [online] Available at:<http://energy.wilkes.edu/pages/184.asp> [Accessed 10/06/2015]

Stamford, L. and Azapagic, A. (2014). Life cycle environmental impacts of UK shale gas. Applied Energy. 134, 506-518.

States, S., Cyprych, G., Stoner, M., Wydra, F., Kuchta, J., Monnell, J. and Casson, L. (2013). Marcellus Shale drilling and brominated THMs in Pittsburgh, Pa., drinking water. Journal-American Water Works Association. 105(8), E432-E448.

Stedman, R. C., Jacquet, J. B., Filteau, M. R., Willits, F. K., Brasier, K. J. and McLaughlin, D. K. (2012). Environmental reviews and case studies: Marcellus shale gas development and new boomtown research: Views of New York and Pennsylvania residents. Environmental Practice. 14(4), 382-393.

Steinzor, N., Subra, W. and Sumi, L. (2013). Investigating links between shale gas development and health impacts through a community survey project in Pennsylvania. New solutions: a journal of environmental and occupational health policy. 23(1), 55-83.

Steinzor, N., Subra, W. and Sumim, L. (2012). Gas Patch Roulette: How Shale Gas Development Risks Public Health in Pennsylvania. . pp. 51

STUK - Radiation and Nuclear Safety Authority. (2012). Some examples of radiation doses. [online] Available at:<http://www.stuk.fi/sateilyvaara/en_GB/esim_annos/> [Accessed 15/06/2015]

Styles, P. and Baptie, B. (2012). Briefing note on induced seismicity in the UK and its relevance to hydraulic stimulation for exploration for shale gas. pp. 12

Subra, W. (2009). Results of health survey of current and former DISH/Clark, Texas residents. Earthworks. pp. 17

Subra, W. (2010). Community health survey results, Pavillion, Wyoming residents. Earthworks. pp. 13 Talabani, S., Atlas, B., Al-Khatiri, M. B. and Islam, M. R. (2000). An alternate approach to downhole

corrosion mitigation. Journal of Petroleum Science and Engineering. 26(1-4), 41-48. Texas Department of State Health Services. (2010a). DISH, Texas exposure investigation Dish, Denton

County, Texas. pp. 38 Texas Department of State Health Services. (2010b). Summary of investigation into the occurrence of

cancer zip codes 75022 and 75028, Flower Mound, Denton County, Texas 1998–2007, 2007–2009. pp. 17

150

Texas Department of State Health Services. (2011). Updated summary report of Texas Department of State Health Services investigation of specific cancer occurrences within zip codes 75022 an 75028, Flower Mound, Denton County, Texas covering 1999–2008, 2008–2010. pp. 17

The Groundwater Foundation. (2015). Potential Threats to Our Groundwater. [online] Available at:<http://www.groundwater.org/get-informed/groundwater/contamination.html> [Accessed 10/03/2015]

The Royal Society and the Royal Academy of Engineering. (2012). Shale gas extraction in the UK: a review of hydraulic fracturing. pp. 76

United Kingdom Onshore Operators Group. (2013a). Fact Sheet: Development and the Drilling Process. [online] Available at:<http://www.ukoog.org.uk/onshore-extraction/stages-of-development> [Accessed 06/01/2015]

United Kingdom Onshore Operators Group. (2013b). Water UK and UKOOG to work together to minimise the impact of shale gas development on water resources in the UK. [online] Available at:<http://www.ukoog.org.uk/about-ukoog/press-releases/63-water-uk-and-ukoog-to-work-together-to-minimise-the-impact-of-shale-gas-development-on-water-resources-in-the-uk> [Accessed 02/05/2015]

United Kingdom Onshore Operators Group. (2015). UK Onshore Shale Gas Well Guidelines: Exploration and appraisal phase. pp. 37

United States Department of Labor Occupational Safety & Health Administration. (2012). Worker Exposure to Silica during Hydraulic Fracturing. [online] Available at:<https://www.osha.gov/dts/hazardalerts/hydraulic_frac_hazard_alert.html.> [Accessed 10/04/2015]

United States Environmental Protection Agency. (2012). An Introduction to Indoor Air Quality (IAQ); Volatile Organic Compounds (VOCs). [online] Available at:<http://www.epa.gov/iaq/voc.html> [Accessed 02/02/2015]

United States Geological Survey. (2009). Monitoring Volcano Ground Deformation with Tiltmeters. [online] Available at:<http://volcanoes.usgs.gov/activity/methods/deformation/tilt/> [Accessed 19/06/2015]

United States Geological Survey. (2014). Monitoring Instruments. [online] Available at:<http://earthquake.usgs.gov/monitoring/deformation/data/instruments.php> [Accessed 19/06/15]

University of Maryland Institute for Applied Environmental Health School of Public Health. (2014). Potential Public Health Impacts of Natural Gas Development and Production in the Marcellus Shale in Western Maryland. [online] Available at:<http://www.marcellushealth.org/uploads/2/4/0/8/24086586/final_report_08.15.2014.pdf> [Accessed 19/04/2015]

University of Michigan Graham Sustainability Institute. (2013). Hydraulic Fracturing in Michigan Integrated Assessment technical reports. [online] Available at:<http://graham.umich.edu/knowledge/pubs> [Accessed 19/04/2015]

US Environmental Protection Agency. (2012). Alpha Particles. [online] Available at:<http://www.epa.gov/radiation/understand/alpha.html#emitters> [Accessed 21/06/2015]

Vengosh, A., Jackson, R. B., Warner, N., Darrah, T. H. and Kondash, A. (2014). A critical review of the risks to water resources from unconventional shale gas development and hydraulic fracturing in the United States. Environmental Science & Technology. 48(15), 8334-8348.

Vidic, R. D., Brantley, S. L., Vandenbossche, J. M., Yoxtheimer, D. and Abad, J. D. (2013). Impact of shale gas development on regional water quality. Science. 340(6134), 1235009.

Voltz, C. D., Ferrar, K., Michanowicz, D., Christen, C., Kearney, S., Kelso, M. and Malone, S. (2011). Contaminant Characterization of Effluent from Pennsylvania Brine Treatment Inc., Josephine Facility Being Released into Blacklick Creek, Indiana County, Pennsylvania: Implications for Disposal of Oil and Gas Flowback Fluids from Brine Treatment Plants [online] Available

151

at:<http://www2.epa.gov/sites/production/files/documents/contaminantcharacterizationofeffluent.pdf> [Accessed 03/05/2015]

Ward, R. S., Stuart, M. E. and Bloomfield, J. P. (2014). Chapter 5: The Hydrogeological Aspects of Shale Gas Extraction in the UK. In: Hester, R. E. & Harrison, R. M. (Eds). Fracking. Royal Society of Chemistry Publishing.

Warner, N. R., Christie, C. A., Jackson, R. B. and Vengosh, A. (2013). Impacts of shale gas wastewater disposal on water quality in western Pennsylvania. Environmental Science & Technology. 47(20), 11849-11857.

Warner, N. R., Darrah, T. H., Jackson, R. B., Millot, R., Kloppmann, W. and Vengosh, A. (2014). New tracers identify hydraulic fracturing fluids and accidental releases from oil and gas operations. Environmental Science & Technology. 48(21), 12552-12560.

Warner, N. R., Jackson, R. B., Darrah, T. H., Osborn, S. G., Down, A., Zhao, K., White, A. and Vengosh, A. (2012). Geochemical evidence for possible natural migration of Marcellus Formation brine to shallow aquifers in Pennsylvania. Proceedings of the National Academy of Sciences. 109(30), 11961-11966.

Watson, S. J., Jones, A. L., Oatway, W. B. and Hughes, J. S. (2005). Ionising Radiation Exposure of the UK Population: 2005 Review. HPA-RPD-001. Health Protection Agency: Chilton. [online] Available [Accessed 02/05/2015]

Weinhold, B. (2012). The Future of Fracking: New Rules Target Air Emissions for Cleaner Natural Gas Production. Environmental Health Perspectives. 120(7), a272-a279.

Werner, A. K., Vink, S., Watt, K. and Jagals, P. (2015). Environmental health impacts of unconventional natural gas development: A review of the current strength of evidence. Science of The Total Environment. 505, 1127-1141.

West Virginia Department of Environmental Protection. (2013). Noise, Light, Dust, and Volatile Organic Compounds Generated by the Drilling of Horizontal Wells Related to the Well Location Restriction Regarding Occupied Dwelling Structures. Office of Oil and Gas publication pursuant to WV code §22-6A-12(e). [online] Available at:<http://www.dep.wv.gov/oil-and-gas/Horizontal-Permits/legislativestudies/Documents/FINAL%20OOG%20Noise%20Light%20Dust%20and%20VOCs%20Report%205-28-2013.pdf> [Accessed 10/04/2015]

Wheeler, D., Atherton, F., Bradfield, M., Christmas, K., Dalton, S., Dusseault, M., Gagnon, G., MacIntosh, C., Mauro, I. and Ritcey, R. (2014). Report of the Nova Scotia Independent Panel on Hydraulic Fracturing. [online] Available at:<http://energy.novascotia.ca/sites/default/files/Report%20of%20the%20Nova%20Scotia%20Independent%20Panel%20on%20Hydraulic%20Fracturing.pdf> [Accessed 19/04/2015]

Williams, D. D., Williams, N. E. and Cao, Y. (2000). Road salt contamination of groundwater in a major metropolitan area and development of a biological index to monitor its impact. Water Research. 34(1), 127-138.

Williamson, D. (2013). Defining drilling fluids: Drilling fluid basics. Oilfield Review. 25(1), 63-64. Wilson, J. M. and Van Briesen, J. M. (2013). Source water changes and energy extraction activities in

the Monongahela River, 2009-2012. Environmental Science & Technology. 47(21), 12575-12582.

Witter, R., McKenzie, L., Towle, M., Stinson, K., Scott, K., Newman, L. and Adgate, J. (2010). Health Impact Assessment for Battlement Mesa, Garfield County, Colorado. Colorado School of Public Health, University of Colorado Denver, Colorado School of Public Health. [online] Available at:<http://www.garfield-county.com/public-health/documents/1%20%20%20Complete%20HIA%20without%20Appendix%20D.pdf> [Accessed 10/04/2015]

Witter, R., Stinson, K., Sackett, H., Putter, S., Kinney, G., Teitelbaum, D. and Newman, L. (2008a). Potential exposure-related human health effects of oil and gas development: A literature review. Colorado School of Public Health. pp. 32

152

Witter, R., Stinson, K., Sackett, H., Putter, S., Kinney, G., Teitelbaum, D. and Newman, L. (2008b). Potential exposure-related human health effects of oil and gas development: A white paper. Colorado School of Public Health. pp. 46

Witter, R. Z., McKenzie, L., Stinson, K. E., Scott, K., Newman, L. S. and Adgate, J. (2013). The use of health impact assessment for a community undergoing natural gas development. American Journal of Public Health. 103(6), 1002-1010.

Witter, R. Z., Tenney, L., Clark, S. and Newman, L. S. (2014). Occupational exposures in the oil and gas extraction industry: state of the science and research recommendations. American Journal of Industrial Medicine. 57(7), 847-856.

Wood, R., Gilbert, P., Sharmina, M., Anderson, K., Footitt, A., Glynn, S. and Nicholls, F. (2011). Shale gas: a provisional assessment of climate change and environmental impacts: A report comissioned by The Cooperative and undertaken by The Tyndall Center, University of Manchester. pp. 82

World Health Organisation. (2015a). Children's environmental health - Endocrine Disrupting Chemicals (EDCs). [online] Available at:<http://www.who.int/ceh/risks/cehemerging2/en/> [Accessed 15/06/2015]

World Health Organisation. (2015b). Radon and health. [online] Available at:<http://www.who.int/mediacentre/factsheets/fs291/en/> [Accessed 15/06/2015]

Younger, P. L., Banwart, S. A. and Hedin, R. S. (2002). Mine water: hydrology, pollution, remediation. 1st ed. Kluwer Academic Publishers, Dordrect.

Zielinska, B., Fujita, E. and Campbell, D. (2011). Monitoring of emissions from Barnett Shale natural gas production facilities for population exposure assessment. Desert Research Institute. pp. 71

Zoback, M., Kitasei, S. and Copithorne, B. (2010). Addressing the environmental risks from shale gas development: briefing paper 1. pp. 18

153

Appendix 1

Appendix 1.1. Table displaying the chemical composition of API Class A, B, C, G and H cements.

154

Appendix 1.2. Table displaying the physical requirements for API Classes A, B, C, G and H cements.

155

Appendix 1.3. Table displaying the some of the physical properties of API Classes A, C, G and H cements.

156

Appendix 2

157

Table displaying the composition of the hydraulic fracturing fluid used by Cuadrilla at the Preese Hall site.