Sei sulla pagina 1di 18

Engineering Geology 168 (2014) 149–166

Contents lists available at ScienceDirect

Engineering Geology
journal homepage: www.elsevier.com/locate/enggeo

Bow-tie risk assessment combining causes and effects applied to gas oil
storage in an abandoned salt cavern
Karin van Thienen-Visser a,⁎, Dimmie Hendriks b, Annemieke Marsman b, Manuel Nepveu a,
Remco Groenenberg c, Ton Wildenborg a, Hans van Duijne b, Marinus den Hartogh c, Tobias Pinkse c
a
TNO, Princetonlaan 6, P.O. 80015, 3508 TA Utrecht, The Netherlands
b
Deltares, Unit Subsurface and Groundwater Systems, Dept. Soil and Groundwater Systems, Princetonlaan 6, P.O. 85467, 3508 AL Utrecht, The Netherlands
c
Akzo Nobel Industrial Chemicals BV, PO Box 25, 7550 GC Hengelo, OV, The Netherlands

a r t i c l e i n f o a b s t r a c t

Article history: A semi-quantitative risk assessment is presented for the storage of gas oil in depleted salt caverns in the Twente
Received 21 June 2012 region, the Netherlands. It is based on a bow-tie model, in which an incident, leakage of gas oil from the storage
Received in revised form 12 July 2013 system (cavern and wells), is evaluated by assessing its possible causes and effects. The causes are all the events
Accepted 10 November 2013
that may lead to leakage from the storage system. The effects are the consequences of the leakage. It is considered
Available online 18 November 2013
that the most serious of the subsurface risks is contamination of the groundwater due to upward migration of the
Keywords:
gas oil to the surface. A unique aspect of our risk assessment is the combination of causes and effects.
Risk analysis The effects of containment/failure are quantified at multiple time scales using a numerical flow model for multi-
Markov Chain phase flow through porous medium, based on the geohydrological properties of the subsurface of the Twente
Storage area. The probability of occurrence of loss of containment/failure (causes) is quantified semi-quantitatively,
Salt caverns using the causal relationships between the causes and effects.
Multiphase flow modelling Modelling of the leakage shows that, as expected, leakage from the well above the hydrogeological base in the
phreatic aquifer produces an immediate risk of contamination of the upper groundwater. However, leakage at
a deeper level does not pose a risk of contamination of groundwater, because the low porosity and permeability
of the geological layers prevent the upward migration of leaking gas oil. The semi-quantitative approach to the
probability of failure finds that for multiple scenarios (e.g. well failure, unstable cavern, high pressure) and in
the absence of human intervention, the probabilities of failure are medium to high. If human intervention is
assumed, these probabilities of failure diminish considerably, especially those associated with the well. These
findings are consistent with those from other hazard studies on storage in salt caverns.
The causes (probabilities of failure) and effects (modelling of leakage) together indicate that for most scenarios
the risk is low when human intervention (e.g. monitoring of the well) is assumed.
Notwithstanding our conclusion that the risk of leakage associated with gas oil injection and storage in salt
caverns is low, an extensive monitoring plan should be formulated to monitor the containment of the gas oil
in the storage system and its long-term stability, to ensure timely human intervention that reduces the risk
considerably.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction formation, which on abandonment are left filled with residual brine
and insoluble materials that together make up the cavern sump area.
Salt solution has been mined since 1919 in the Twente region of For reasons of ground stability and to avoid subsidence and collapse,
eastern Netherlands from horizontal thinly bedded salt deposits at caverns have a permitted spacing (Figure 3); in a large proportion of
relatively shallow depth (300–500 m). In salt solution mining, fresh or the salt caverns in the Twente area, salt reserves have been depleted
undersaturated water is injected at the top of the cavern through the and production has ceased. A new use for these depleted salt caverns
annular space between the cemented casing and the brine production is the storage of hydrocarbons (crude oil, gas oil). Storage is achieved
string, while brine from the dissolution of the salt beds is produced by withdrawing some of the residual brine from the bottom of the
from the bottom of the cavern through the production string. The cavern while simultaneously injecting hydrocarbons at the top of the
Twente caverns have been produced from the bottom up. The salt cavern. During retrieval this process is reversed.
solution mining process creates caverns (voids) within the rock salt Storage of hydrocarbons in salt caverns is applied in many countries
around the world (Thoms and Gehle, 2000). The earliest cavern storage
⁎ Corresponding author. Tel.: +31 88 86 64265; fax: +31 88 86 64475. in salt domes for liquefied petroleum gas (LPG) started in 1951 (Avery
E-mail address: karin.vanthienen@tno.nl (K. van Thienen-Visser). Island, USA); LPG storage in bedded salt deposits started earlier, in the

0013-7952/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.enggeo.2013.11.002
150 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

early 1940s (USA, Querio, 1980). In Europe, hydrocarbon storage in of the cavern and wells (including over/underburden and faulting),
depleted salt caverns has been practised since the early 1950s (Thoms the long-term stability of the caverns, and possible effects on the envi-
and Gehle, 2000). Examples in Europe include the Manosque facility ronment if leakage occurs. For this risk assessment we use the concept
in France and the Etzel salt dome near Wilhemshaven, Germany, of a bow-tie (Figure 2). Central in this bow-tie model is an incident
which has been used since 1971 for crude oil storage. Salt cavern storage that has a certain probability of occurrence. Each incident has different
is one of the safest methods of storing large amounts of hydrocarbons causes and effects, which are indicated to the left and right of the
(Bérest et al., 2001; Bérest and Brouard, 2003), mainly due to the bow-tie, respectively. The causes (left side of bow-tie, Figure 2) are
impermeability of the salt. Nonetheless, incidents have occurred at 9% events that may lead to loss of containment and the effects (right side
of the salt cavern storage facilities worldwide (Evans, 2008; Evans and of bow-tie, Figure 2) are the consequences of the loss of containment
Chadwick, 2009). Almost all incidents are due to human error and for the contamination of groundwater (harmful effect). Causal relations
poor safety controls and checks. The reported incidents can be divided that exist between the causes and effects of each incident can be
into design failure (34%), operational failure (30%), well failure (26%) translated into hazard scenarios. To prevent the incident central to the
and failure of surface infrastructure (8%). In some incidents, failure bow-tie, measures can be taken on both sides of the bow-tie. Causes
resulted from a combination of problems. can be prevented by dedicated monitoring and corrective measures,
Prior to selecting caverns for the storage of gas oil (liquid fuel), it is and remaining effects can be mitigated.
necessary to screen potential storage caverns for suitability. The most The incident considered is leakage of gas oil from the storage system
important criterion is that gas oil cannot leak from the storage system (loss of containment). The storage system consists of abandoned salt
into groundwater and aquifers and thence to the surface, i.e., the risk caverns in Twente (the eastern part of the Netherlands). Their (hydro)
of groundwater contamination and human safety must be reduced to geological setting is described in Section 2. In Section 3, the causes are
near zero. Therefore we introduce the Containment Concept (Figure 1). assessed in the light of the literature; the effects are modelled in
The Containment Concept for subsurface storage encompasses the entire Section 4, using a multiphase flow model. Finally, in Section 5 we
suite of barriers and facilities that ensure that the gas oil stored in a salt consider the bow-tie as a whole and quantify the total underground
cavern does not escape the storage caverns and disperse into the risk using expert elicitation (Wildenborg et al., 2005; Knol et al., 2010).
surrounding storage environment/system. The boundary of the con- Mitigation of risk associated with the aboveground part of the oil storage
tainment system for the gas oil is formed by the cavern walls, the gas system (e.g. transport of gas oil) leading to risk to human safety is not
oil–brine interface at the base of the cavern, the cavern roof, and the considered here but has been addressed in the Environmental Impact
wells (packers, casing, and wellhead) and is represented by the green Assessment (Tauw, 2013).
line in Fig. 1. In the case of a working containment, it is unlikely that
the gas oil will penetrate the surrounding salt for more than several 2. Natural containment: the hydrogeological setting
millimetres in a thousand years, due to the very low permeability of
the salt (Bouw and Oude Essink, 2003; Doe and Osnes, 2006). The hydrogeology of the Twente area consists of alternating aquifers
Considering a possible failure of containment, an assessment of the and aquitards and provides the natural containment of gas oil in the salt
risks associated with gas oil (liquid fuel) storage in depleted salt caverns caverns. Key parameters controlling the flow of liquids through the
in Twente is required. In this paper, we present a risk assessment subsurface are the permeability of the different stratigraphic units, the
methodology that takes into account the integrity of the containment presence of faults, and the density of the groundwater. Finally, the

Fig. 1. Illustration of the Containment Concept applied in the risk assessment study for the storage of gas oil in abandoned salt caverns in the Twente area, the Netherlands.
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 151

Fig. 2. Bow-tie model used for the risk assessment.

position and dimensions of the caverns in the salt layers are important contains the salt caverns (NITG-TNO, 1998; Geowulf, 2008; MWH,
for containment. 2010). The member consists primarily of halite, but has a thick anhy-
drite layer at the base, as well as intercalated clay layers of 10 to 15 m
2.1. Stratigraphy of the Twente area thickness at the top. It is subdivided into four units by the presence of
four claystone intervals (with variable dolomite and anhydrite) that
The stratigraphy in the Twente area from deep to shallow (old to are well correlated over the general area. In the Twente area, the four
young) comprises the Solling Formation, the Röt Formation (containing salt layers have been named, from top to bottom, A, B, C and D. Salt
the salt caverns), the Muschelkalk Formation, the Altena Group, the layer A has a thickness that varies between 10 and 60 m, while layer C
Niedersachsen Group and the North Sea Supergroup. As a result of has a more constant thickness of 15 to 20 m. Salt layers B and D are
tectonic activity, there are several fault zones in the region (NITG- generally in the order of only several metres thick. More information
TNO, 1998). The Röt Formation, where the salt caverns are located, is about the lithology and thickness of all geological layers can be found
subdivided into the Main Röt Evaporite Member, the Middle Röt in Table 1 (NITG-TNO, 1998; Geowulf, 2008; MWH, 2010).
Claystone Member, the Upper Röt Evaporite Member and the Upper
Röt Claystone Member. The thickness of the Röt Formation in Twente 2.2. Hydrogeological properties of the subsurface
varies from 225 m in the north to 300 m in the centre of the area and
decreases to slightly less than 200 m in the south. The Main Röt There are reported values for porosity and permeability or hydraulic
Evaporite Member, which is the oldest member of the formation, conductivity of the geological formations that occur in the subsurface of

Table 1
Hydrogeological and hydraulic properties of the geological units of the conceptual model of the Twente area. Source references of the hydraulic properties are: 1. Verweij and Simmelink
(2002), 2. Bouw and Oude Essink (2003), 3. Pöppelreiter et al. (2005), 4. Doe and Osnes (2006), 5. Bear (1972), 6. Domenico and Schwartz (1998), Dufour (1998), De Louw (2006).
Schematisation for STOMP modelling is also given: layering within the geological formations and the layer thickness.

Description host rock Properties host rock Properties permeable faults

Geological Formation Lithology STOMP STOMP layer Particle Total Geological Hydraulic Total Geological Hydraulic
formation thickness depth top thickness density porosity permeability conductivity porosity permeability conductivity
Source hydraulical (m) layer (m) (kg/m3) (fraction) (mD) (m/d) (fraction) (mD) (m/d)
properties (m)

North Sea 10–60 Sand (silt, clay) 0 28 2082 0.41 1.00E+05 1.00E + 00 – – –
Supergroup
above hydrogeol.
base 6,7,8
North Sea Clay −28 10 1746 0.1 1.00E−02 1.00E−08 – – –
Supergroup Consolidated −38 10 2200 0.3 5.00E+02 5.00E−04 – – –
below hydrogeol. sand
base 1,6,58 100–150 Claystone −48 10 1746 0.1 1.00E−02 1.00E−08 – – –
Consolidated −58 10 2200 0.29 5.00E+02 5.00E−04 – – –
sand
Claystone −68 10 1746 0.1 1.00E−02 1.00E−08 – – –
Altena group/ 0–100 Claystone −78 40 1746 0.07 1.00E−02 1.00E−08 0.17 1.00E−01 1.00E−07
Niedersachsen
group1,2,56
Muschelkalk3,5,6 Dolomitic marl −118 20 2243 0.27 1.00E+00 1.00E−06 0.37 1.00E+02 1.00E−04
Clayey marl −138 20 2243 0.29 1.00E−02 1.00E−08 0.39 5.00E−01 5.00E−07
0–200 Dolomitic marl −158 30 2243 0.27 1.00E+00 1.00E−06 0.37 1.00E+02 1.00E−04
Clayey limestone −188 20 2243 0.14 1.00E+01 1.00E−05 0.24 1.00E+03 1.00E−03
Marl, anhydrites −208 20 2610 0.24 1.00E−01 1.00E−07 0.34 1.00E+01 1.00E−05
Upper Röt 135–200 Claystone −228 170 1860 0.07 1.00E−02 1.00E−08 0.17 1.00E−01 1.00E−07
Claystone2,5,6
Upper Röt 15 Anhydrite −398 15 2320 0.01 1.00E−05 1.00E−10 0.11 1.00E−04 1.00E−09
Evaporite2,4
Main Röt 20–110 Rock salt, −413 60 2150 0.005 1.00E−05 1.00E−10 0.005 1.00E−05 1.00E−10
Evaporite2,4 claystone
152 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

the Twente area (see Bear, 1972; Domenico and Schwartz, 1998; and 1500 m (Jurassic and Cretaceous). At such depths, the high pres-
Dufour, 1998; Verweij and Simmelink, 2002; Bouw and Oude Essink, sures and temperatures intensify the fault-healing process in the salt.
2003; Pöppelreiter et al., 2005; De Louw, 2006; Doe and Osnes, 2006). Additionally, fluid inclusions in the original salt stimulate pressure
Although these values have been obtained partly from locations outside solution and grain boundary migration recrystallization, with the result
the Twente area, they give a good indication of the porosity and perme- that the offsets of the faults are accommodated by changes in crystal
ability of the subsurface of the Twente area. structure instead of by brittle failure. Since the rock salt layer of the
Hydrogeologically, the subsurface of the Twente area can be Main Röt Evaporite which contains the salt caverns is at least 25 m
subdivided into two domains: the permeable aquifers in the sandy thick, the maximum fault throw of 10 m that has been reported for
deposits above the hydrogeological base and the generally less perme- the area is too small to achieve complete offset of the bed (Geowulf,
able or impermeable deposits below this hydrogeological base. The 2010).
first 10 to 60 m of the subsurface consist mainly of unconsolidated
sandy deposits with a high permeability (1.00 × 101–1.00 × 106 mD).
They constitute the phreatic aquifer that is important for abstraction 2.4. Groundwater density
of drinking water, water for sprinkling and irrigation, aquatic and
terrestrial ecology and river and stream runoff (Hendriks et al., The density of the groundwater greatly influences the behaviour of
accepted). Depth to groundwater in the area ranges between 0 and gas oil in the shallow subsurface (see also Section 5). Groundwater
more than 12 m; and the groundwater flow in the area is from the topo- density mainly depends on how much dissolved salts the water con-
graphically high areas (ice-pushed ridges) to the low areas (De Louw, tains: chloride and sulphate concentrations are a proxy for this. Because
2006; Hendriks et al., accepted). fresh water is less dense than salt and brackish water, it tends to float on
At the base of this shallow aquifer lie marine clays that have very low the saline groundwater in Twente. Only limited mixing of the fresh and
permeability (1.00 × 10−5–1.00 × 101 mD). They form a significant saline groundwater types occurs. In the Twente area, there is a transi-
barrier (the hydrogeological base) and reduce connectivity with deeper tion from fresh to salt groundwater in the Tertiary deposits of the
aquifers (De Louw, 2006). All the geological layers between the Upper North Sea Group. The Peize Formation and younger, shallower
hydrogeological base and the Main Röt Evaporite in which the salt formations were deposited in a continental environment, whereas the
caverns are located have high resistances to flow (1.00 × 101– older formations were deposited in a marine environment. However,
1.00 × 106 mD). However, parts of the Muschelkalk Formation and as a result of the precipitation that has fallen on the Twente area since
the North Sea Super Group consist of aquifers (NITG-TNO, 1998). the deposition of these formations, fresh water has infiltrated towards
In our study, the values used for the porosity and permeability of the deeper aquifers and the original freshwater boundary has shifted down-
salt layers of the Main Röt Evaporite, in which the salt caverns are ward. Consequently, the groundwater above the hydrogeological base
situated, were typical values for rock salt and anhydrite reported in the consists of fresh water, and below it the groundwater is brackish or
literature. For rock salt and anhydrite, Doe and Osnes (2006) reported salt (chloride concentrations of 1500 to 15000 mg l−1 and sulphate
a hydraulic conductivity of 7.0 × 10−8 to 1.6 × 10−5 m d−1, and concentrations of 250 to 1400 mg l−1). In the younger Quaternary
Bouw and Oude Essink (2003) assign a maximum hydraulic conductivity deposits the groundwater quality varies little: chloride concentrations
of 1.00 × 10−13 m s−1 (1.00 × 10−8 m d−1) and a porosity of 2%. An range from 35 to 50 mg l−1 (Dufour, 1998).
overview of the hydrogeological properties of all geological units in the
subsurface is given in Table 1.
2.5. Cavern position and dimensions
2.3. Permeability of faults
The salt caverns in Twente lie at depths between 300 and 500 m.
In the subsurface of the Twente area, normal faults with a maximum They are located in the Main Röt Evaporite Member (Table 1, Figures 1,
fault throw of 10 m have been detected both above and below the salt 3 and 4), which forms part of the Triassic Röt Formation (NITG-TNO,
(NITG-TNO, 1998; Geowulf, 2010). Some of the faults extend up to, 1998). Because the salt layers are horizontal and thinly bedded, the
but never into, the deposits of the North Sea Group. If fault permeability caverns are relatively small compared to salt caverns in other regions
is strongly anisotropic, faults can form a preferential pathway between (Thoms and Gehle, 2000). They are up to 40 m high and have a long
aquifers at different depths over vertical distances of several hundreds axis of between 100 and 300 m and a short axis of 100 m (Figure 3).
of metres that are otherwise separated by confining units. This has The caverns considered in this study were those situated in the Röt
important implications for the assessment of the risk of a spread of salt layer A for reasons of stability. Caverns extending in Röt layer
contaminated groundwater or the reconstruction of hydrocarbon B and C are generally less stable and therefore not suited for gas oil
migration within sedimentary basins (Bense and Person, 2006). How- storage.
ever, despite various studies (e.g. Bense and Person, 2006; Anderson
and Bakker, 2008; Folch and Mas-Pla, 2008; Magri et al., 2010; Saar,
2010), no explicit assumptions can be made on the width and the 3. Causes of loss of containment
hydraulic properties of permeable faults. Generally, fault width increases
with increasing fault throw (fault width of approximately 2 m with fault Possible causes of breach of containment in the Twente area and the
throw of 50 m; fault width of approximately 4 m with fault throw of associated levels of risk and requirements to prevent them had been
100 m). The permeability of such faults can be assumed to be several identified earlier (van Duijne et al., 2011; Tauw, 2013). The present
orders of magnitude greater than that of the surrounding host rock, study focused on the most likely scenarios (Figure 4) including salt
depending on the clay content of the host rock (Bense and Person, creep and temperature effects. Causes of breach of confinement were
2006). Based on this knowledge, we made conservative estimates of defined as leakage fluxes from the storage system into the surrounding
the fault properties in the Twente area: fault width of 0.2 to 5 m and rock and were separated into two categories: those requiring a detailed
fault permeability of 1.00 × 10−1–1.00 × 102 mD depending on the approach because the risks involved were higher, and those requiring a
host rock lithology. more global approach because the risks involved were lower. The
In salt layers the effect of slippage along normal faults is different. detailed approach entailed a semi-quantitative risk assessment based
Salt is viscoplastic, so it tends to heal fractures in fault zones prohibiting on literature research, knowledge of the specific geology of the region
an increase of permeability due to tectonic activity. The tectonic activity and expert assessment, the more global approach required only a gen-
in the area occurred when the salt was situated at a depth between 500 eral literature research.
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 153

Fig. 3. Geometry of salt production caverns in Twente. The left side of the figure contains details of the two cross-sections shown on the map on the right. Section (A–A’) shows the minimal
permissible width for the salt pillar separating two adjacent caverns in a row that is required to prevent pressure communication. Section (B–B’) shows the minimal permissible width for
the salt pillar between adjacent rows of salt caverns that is necessary in order to maintain stability.

Fig. 4. Schematic overview of the possible leakage fluxes from the storage cavern: flux through the walls (1), flux through the floor (2), flux through the roof (3), flux through the casing
(4a), flux through the casing shoe (4b), and flux through faults or fissures (5).
154 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

3.1. Flux from the cavern fracturing may occur if the pressure in the cavern exceeds the minimum
in situ stress, for instance through overfilling of the cavern. This would
Flux of gas oil from the cavern (through the walls, roof or floor) may be averted by keeping the maximum operational pressures below the
occur through the surrounding rock (in this case, rock salt) or transpor- lithostatic pressure. The lithostatic pressure (pgz, where ρ is density, g
tation through permeable layers or along faults intersecting the cavern is the gravity constant 9.81 m/s2 and z is the depth) ranges from 54 to
or immediate surroundings. If the Containment Concept is working, 90 bar (assuming an average overburden density of 1800 kg/m3 and a
the gas oil can infiltrate into the surrounding salt (Röt A: rock salt) for depth ranging from 300 to 500 m). In the case of a cavern filled with
only a small distance due to the salt's low permeability and hydraulic con- gas oil (the density of gas oil is 850 kg/m3), the pressure increase inside
ductivity (k = 1.16 × 10−2 mD; K = 7.0 × 10−8 to 1.6 × 10−5 m/d) the cavern is at most 1.4 bar (pgz = 1.4 bar, assuming the cavern is
(Doe and Osnes, 2006). Since the caverns are completely enclosed by 40 m high and the difference in density between the gas oil and brine
rock salt (Figure 4) with a safety buffer of at least 25 m between adja- is 350 kg/m3). This is too small to trigger tensile fracturing.
cent caverns in a row, to prevent pressure communication (BGR, A cavern might be unstable because its salt roof is too thin to cope
2004a), a distance of 70 m between adjacent rows of salt caverns in with the pressure differences (stresses) acting on the cavern or
order to maintain stability and a 5 m thick rock salt roof in Röt salt A, resulting from another cavern that is too close by. Using the thicknesses
it is expected that the gas oil will not infiltrate into the surrounding and properties of the geological layering in Twente, BGR (1998) per-
rock during the lifespan of the storage (30 years). Assuming conductiv- formed a geomechanical modelling study and calculated that to ensure
ity is high, the maximum distance the gas oil will infiltrate the stability, the minimum thickness of the salt roof in the Röt salt C must be
surrounding salt is 0.18 m (1.6 × 10−5 m/d × 365 days × 30 years). 5 m. Also important for stability is the shape of the cavern roof. Prior to
A special case is the cavern floor, as it does not come into contact with gas oil storage, the roof will be hollowed out into a bell-shaped dome
the gas oil under normal operating conditions due to the residue of around the well. This has an added advantage that gas oil will be easier
brine and insoluble material that collects in the cavern sump area. A to retrieve at the end of storage. The thickness of the salt roof is per se
variable volume of brine (occupying 5–20% of the cavern) will remain not vital to the stability of the cavern roof area. What determines the
in the cavern due to the height of the injection well inside the cavern; stability of the cavern roof is the thickness of the salt roof combined
floating on top of it will be a layer of gas oil, which at 850 kg/m3 is with the properties of the geological layers above the cavern and the
less dense than the brine (density 1200 kg m−3). As the caverns were presence of intersecting faults or faults in the immediate surroundings,
solution mined from the base upwards, if the pressure inside the cavern as well as the internal hydrostatic pressure (brine static pressure).
is much lower than the hydrostatic pressure, an influx of water could
occur from the Solling Formation below the cavern (Van Sambeek 3.3. Leakage of wellbore
et al., 2005). If there has been additional dissolution, the inner cavity
pressure will decrease with time before the injection of gas oil. This Leakage through the wellbore has been indicated as a major risk in
scenario is, however, unlikely, since it would require pressures much the storage of gas oil in salt caverns (Evans, 2008; Evans and Chadwick,
smaller than the planned operating pressure during storage, which is 2009). An area requiring particular attention is that of the placing of
32 bars. the casing shoe and its cementation in the cavern roof area. Several
Considering the extremely low permeability of the lithology in the studies report leakage through this section of the well during hydrocar-
immediate vicinity of the caverns (rock salt, clays), leakage through bon storage activities (e.g. Evans, 2008; Evans and Chadwick, 2009). In
the cavern walls is considered unlikely under normal operating condi- the context of well integrity the interfaces between different materials
tions, in which pressure changes are planned to be minimal. The hydro- (cement/rock, cement/casing, casing/plug) have been identified as
static pressure due to the brine (density is 1200 kg/m3) in the cavern potential preferential pathways for fluid flow (Gasda et al., 2004;
varies, depending on the depth. It ranges between 35.3 and 58.9 bar Miyazaki, 2009). In the cement annulus, migration of fluid or gas
in the depth range from 300 to 500 m. As long as the hydrostatic column under pressure may occur through pore bonding, fractures, channels,
in the well is determined by the brine, conditions in the gas oil storage or through the pore space. Fluid flow through the pore space in the ce-
phase are the same as in the salt solution mining phase. In the special ment will occur only if the cement is degraded or did not form properly
case of salt caverns adjacent to each other, stress gradients due to during the emplacement process (Zhang and Bachu, 2011). Fluid flow
different pressures inside neighbouring caverns can be prevented and degradation rates in intact cement are too slow to compromise
from building up in the salt pillars between the rows of caverns by well integrity at conceivable time scales (e.g. Kutchko et al., 2007).
ensuring that there is a minimum thickness of 70 m rock salt between The actual state of the cement and the casing are directly related to
adjacent rows of caverns (BGR, 2004b, Fig. 3). This safety buffer must the design of the well and its operational history. Casing displacement
be maintained during the lifetime of the storage site until all facilities and stress changes may cause the cement to fail under tension or
have been properly abandoned. Ideally, caverns used for gas oil storage compression, or to de-bond from the casing or formation, creating
should be surrounded by abandoned caverns to avoid problems arising micro annuli and radial cracks (Boukhelifa et al., 2004). The packer
from adjacent operational brine caverns (for example, over-mining isolates reservoir fluids from the production casing, and “forces” the
which threatens the safety buffer). fluids into the tubing. In addition, the packer may bear some of the
Another possible pathway for fluxes is permeable faults. In Twente tubing loads (depending on how the completion is set). Like the pro-
area several faults have been identified, mostly in and around the duction tubing, evidence for failure of the packer is almost always
Boekelo fault zone which runs NNW–SSE along the southern edge of directly observed (failure of pressure test during initial installation,
the concession area (GeoWulf, 2008; MWH, 2010). As discussed earlier, loss of annulus fluid levels, presence of reservoir fluids inside produc-
the increase in permeability resulting from the presence of the faults tion casing during production life, pressure communication between
will probably be small due to the small fault throws, the healing nature production tubing and production casing). Leakage through packers
of the salt and clay smear in the faults. has rarely led to accidents. Out of 36 recorded accidents (Bérest and
Brouard, 2003; Evans, 2008) in salt caverns, only one involved a packer.
3.2. Flux through the cavern roof The packer failed during repair of the casing and not due to corrosion or
mechanical strain. This indicates that packers usually perform well in
A specific case in which the penetration of gas oil into the roof of the hydrocarbon storage operations. Nevertheless, packers are mechanical
cavern may occur is leakage via cracks formed in response to stress and require special attention as they present the first and direct barrier
changes. Cracks could form in the cavern roof in the case of tensile between the cavern and potential leakage pathways along or out of the
(i.e. extensional) or shear (i.e. compressional) fracturing. Tensile wellbore.
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 155

3.4. Effect of salt creep such fluctuations may occur if the gas oil or brine injected is cold
because the aboveground temperature is low (winter). Significant
A general equation for deformation by creep is changes in creep velocity are, therefore, not expected and will not be
an issue, providing that the pillars separating the caverns are not eroded
dε −Q
n due to the injection of undersaturated brine or the increase in solubility
¼ Ae RT σ ; ð5Þ
dt as a result of rising temperature. Using the IUPAC-NIST Solubility
Database, a rise in temperature from 0 to 20 °C may cause an additional
where ε is the creep strain, A is the proportionality constant (MPa−n s−1),
2.4 g of salt to dissolve per litre of water. Assuming a cavern containing
Q is the activation energy (J/mol), R is the gas constant (J/mol/K), T is the
80% gas oil and 20% brine, the volume of the gas oil, for a 250,000 m3 salt
absolute temperature, σ is the in situ stress and n is the stress exponent
cavern is 200,000 m3 or 2.0 · 108 l. In the case where the gas oil is
related to the micro-mechanism of deformation. Even from the general
removed at the end of a storage cycle and relatively cold brine is injected
salt creep equation, it is clear that creep is highly dependent on temper-
into the cavern, the 80% of gas oil is replaced by 80% relatively cold brine
ature, stress and type of material. More stress/higher temperature will
which will gradually heat up to geothermal temperatures. For a temper-
accelerate creep.
ature increase of 20 °C, 2.4 g more salt may dissolve per litre, resulting
Since the geothermal temperature increases with depth, creep also
in 2.4 · 105 kg extra dissolved salt. For the 80% relatively cold brine
depends on depth.
this translates to 111 m3 more salt that may dissolve (assuming a
For a spherical cavern, the volumetric creep rate is defined as (Van
density of salt of 2165 kg/m3), which corresponds to a cavern volume
Sambeek et al., 2005)
increase of 0.05%. The results of this calculation may vary if a different
. . solubility is assumed: in this case, there may be less erosion of the
dV dR  n
1 3 −Q
cavern walls. In addition, the figure of 80% gas oil being replaced by
dt
¼ dt
¼− ðP ∞ −P i Þ Ae ℜT : ð6Þ
3V R 2 2n brine is high. More accurate results can be obtained by running simula-
tion models of the actual simulation. When the pressure on the gas oil is
The cylindrical case can be deduced from the spherical case by reduced for transportation, the light components of the gas oil are
 pffiffiffinþ1
released until the vapour pressure of the warmed oil is atmospheric.
multiplying by 2= 3 . In a typical storage cavern in Twente, the
This could result in atmospheric discharges of volatile organic
depth will be 300 m and temperature will be 20 °C, resulting in compounds and methane that exceed permitted thresholds, or in safety
sphere cylinder
volumetric creep rates of V̇ =V ¼ −0:0014%=year and V̇ =V ¼ concerns due to local concentrations of flammable gas. The temperature
−0:0026%=year when using Avery Island parameters (Van Sambeek rise in the Twente caverns is very small (b20 °C) because in this area the
et al., 2005). These parameters are used because, like the Rőt salt A salt is at shallow depths (300–500 m), which makes it unlikely that the
layer considered here, Avery Island salt is almost pure halite (99.5%). TVP (true vapour pressure: equilibrium partial pressure exerted by a
Assuming that internal pressure (Pi) in the cavern is 3 MPa lower than volatile organic liquid as a function of temperature) will exceed atmo-
the overburden pressure (P∞), the corresponding volume loss ranges spheric pressure (Giles, 1994). Furthermore, vapour release has not
from 3.6 m3/year in the case of a spherical cavern, to 6.5 m3/year in been observed in similar storage caverns in Wilhelmshaven operated
the case of a cylindrical cavern. In a 30-year period, the total volume by the Nord-West Kavernengesellschaft GmbH (NWKG), Germany.
loss would be 195 m3, which is a small volume loss (0.078%) given the
cavern volume (250,000 m3). 4. Effects of loss of containment
The stress state in the salt around a cavern depends on (1) the
original in situ state of stress, (2) the pressure in the cavern relative Organic fluids such as gas oil pose a serious threat: to drinking
to the lithostatic pressure, (3) cavern geometry, (4) salt creep, and water and ecology, for example. Contamination consisting of 1 l of oil
(5) temperature changes in the salt (Nieland and Ratigan, 2006). There- can significantly affect the quality of 100,000 l of water, i.e. make it
fore, salt creep can be induced by the creation of a void (cavern) in the non-potable for humans. For this reason, the risks of leakage to ground-
salt body, pressure changes in the cavern, stress concentrations around water resources are taken very seriously (Domenico and Schwartz,
the cavern, and stress gradients of the salt pillars separating caverns 1998; Marsman, 2002; Gerritse et al., 2009). Gas oil is an organic Non-
with different pressures. In a typical salt cavern in Twente, given the Aqueous Phase Liquid (NAPL) that is only slightly miscible with water
depth at which these caverns occur (300–500 m), the properties of and has a lower specific density than water. Because its specific density
the Röt salt layers and assuming pressure changes are negligible, the (820 to 860 kg/m3) is lower than that of fresh water (1000 kg m−3) and
velocity of the salt creep is in the order of a few centimetres per year saline water (1025 kg m−3), gas oil “floats” on groundwater. When gas
(van Duijne et al, 2011; personal communication C. Spiers, Utrecht oil leaks from a cavern and no aquitards are present, it migrates upward
University). This velocity is low, especially given that the long axis of towards the water table (Marsman, 2002). To quantify the risk of this
the cavern is from 100 to 300 m (Figure 3). In horizontal thinly bedded happening, first the possible migration paths were identified and corre-
salt deposits, caverns are wide but shallow, and their flattened cross- sponding flow scenarios were defined. Next, the spread of gas oil for
section is therefore inherently associated with stress concentrations at these scenarios was simulated using a numerical flow model suitable
the sides of the cavern, and in the centre of the floor and roof. However, for modelling multiphase flow through porous media (STOMP).
as these stress concentrations were also present during brine produc-
tion, salt creep is not expected to be a high risk factor. 4.1. Migration paths and flow scenarios

3.5. Temperature changes The rate of upward migration of gas oil, the amount of lateral spread,
and the period after which pollution of the phreatic groundwater occurs
Temperature changes in the cavern will induce pressure changes as all depend on the characteristics of the geological layers in the subsur-
well as short term changes in creep velocity and thermal contraction or face around and above the point of leakage (porosity, permeability,
expansion of the salt close to the well. Other effects are increased solu- specific density), as well as on the properties of the gas oil and the
bility of the salt when temperature rises (Kunstman and Urbańczyk, groundwater (density, viscosity, relative permeability). Additionally,
2008) and an increase in temperature of the gas oil due to the geother- the depth of the leakage point is of importance (from the cavern or at
mal gradient (Eyermann, 1994; Giles, 1994) which may cause vapour to a certain depth along the wellbore in the case of failed packers and
be released. Since the caverns are at shallow depths (300–500 m) the faulty casing/cementation). As described above, the overburden of the
maximum fluctuation in temperature will range from 15 °C to 20 °C; Twente area consists of alternating aquifers and aquitards, and contains
156 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

faults in which permeability is higher than in the host rock. The follow- multiphase flow (Marsman, 2002) and on the constitutive relations
ing gas oil migration paths have been defined that might lead to gas oil developed by Lenhard and Parker (1987) and Kaluarachchi and Parker
entering the upper groundwater bodies after loss of containment (5): (1992). Flow and transport are solved numerically using an integrated
volume finite-difference scheme to discretise the governing equations.
(1) Leakage occurs from the well above the hydrogeological base.
For our analysis, the local subsurface of the Twente area was
Leakage causes direct risk of contamination/pollution of the
schematised in a cross-sectional (2D) model domain based on the
shallow groundwater bodies.
hydrogeological information on the Twente area (see Section 3). In
(2) Leakage of gas oil occurs in the well below the hydrogeological
the horizontal plane, the resolution of the model domain is 10 m. Verti-
base. Two scenarios are possible:
cally, the resolution of the model domain varies depending on the level
(a) Breach of containment of the well is local, and the well casing and
of detail that is required in the result and how much is known about the
well shoe are intact. The gas oil will leak into the surrounding
geological layers. The layers that represent the Upper Röt Claystone, the
hard rock or sediment;
Muschelkalk Formation, the Altena Group, the Niedersachsen Group,
(b) Loss of containment of the well is not local and the cement of the
and the North Sea Supergroup below the hydrogeological base have a
well is not intact, creating a short cut for the gas oil towards
resolution of 10 m. The Main Röt Evaporite Member, the Middle Röt
layers above the hydrogeological base from where it will leak
Claystone Member and the Upper Röt Evaporite Member have a much
directly into the upper groundwater.
higher resolution (2 m), as does the layer above the hydrogeological
(3) A permeable fault or another old well with degraded permeable
base (2 m). An overview of the subsurface schematisation of the Twente
cementation is present at the location of the cavern and/or
area is given in Table 1.
the well, so leakage from the cavern or the well below the
For the analyses of the effects of gas oil leakage, a typical salt cavern
hydrogeological base can lead to relatively fast upward move-
was schematised in the STOMP model. It had a total volume of
ment of gas oil towards the upper reach of the fault or old well.
150,000 m3, a height of 35 m, a width of 110 m, and a length of
(4) Gas oil leaking from the cavern or well moves upward towards
160 m. The cavern was assumed to be situated within the Main Röt
shallow groundwater bodies through a combination of permeable
Evaporite, which was 60 m thick in our schematisation. The injection
faults and old wells near the cavern or the well and permeable
well was positioned in the centre of the cavern. Due to the layer of
layers in the overburden.
brine on the floor of the cavern to shield the gas oil from the permeable
Based on these migration paths, we defined seven flow scenarios, all sump material and formations below the cavern into which it may
of which result from loss of containment or a deviation from the base potentially migrate, the maximum volume of gas oil that can be stored
case (see Table 2). For each flow scenario, the depth of the point of was estimated to be 105,000 m3.
leakage was chosen to be within the limits of that flow scenario. Addi- Additionally, the location and the properties of potentially perme-
tionally a maximum period of leakage of 1 to 3 months was defined: able faults and “old” wells with reduced integrity in the vicinity of the
this is assumed to be the period after which mitigation measures start cavern were schematised. In the STOMP model schematisation two
and the leakage is stopped (see Table 3). The maximum period of leak- normal faults at distances of 35 m and 40 m from the central well of
age (before mitigation starts) strongly depends on how monitoring is the salt cavern were schematised as well as an “old” well at a distance
performed. We assumed that cavern pressure is continuously measured of 40 m from the central well of the salt cavern. For permeable faults,
and that within two weeks of observing any pressure irregularity, thor- the permeability can generally be considered to be several orders of mag-
ough investigations are carried out. In the case of leakage from a well, nitude higher than that of the surrounding host rock (see Section 3),
the problem can be solved within two weeks after the investigation. In and depends primarily on the fault throw and the clay content of the
the case of leakage from a cavern, the problem is assumed to be solved rocks flanking the fault zone (Bense and Person, 2006). Because the
within approximately three months. subsurface in the Twente area consists of alternating clay-rich and
sandy formations, the permeability of faults varies strongly with
4.2. Multiphase modelling with STOMP depth. Furthermore, “old” wells in the vicinity of the cavern that may
have a reduced casing integrity, were assumed to have a cement porosity
We used the numerical model STOMP (Subsurface Transport Over of 0.8 and a permeability of 1.00 × 106 mD. The zone of increased per-
Multiple Phases) developed by White et al. (1995), and Lenhard et al. meability ran from 30 m above the roof of the cavern to the surface
(1995) to simulate the spread of gas oil in the subsurface of Twente (worst-case scenario).
when leakage occurs from a storage system with a salt cavern having To simulate of multiphase flow behaviour with the STOMP model,
dimensions typical for the Twente area (see Section 2.5). STOMP is a data must be input on the properties of the subsurface, the formation
three-dimensional, three-phase, compositional engineering simulator water, the brine and the gas oil (Lenhard et al., 1995; White et al.,
for modelling contaminant migration and remediation technologies 1995; Marsman, 2002). Required subsurface data include:
for the clean-up of subsurface sites contaminated with organic com-
pounds. The STOMP model code is based on the Richards equation for • Hydraulic properties of the subsurface: porosity and intrinsic perme-
ability of each defined rock/soil type. Intrinsic permeability can be
Table 2 based on measurements or calculations.
Scenarios of gas oil flow after leakage for the case-specific risk analyses. • Hydraulic conductivity at reference conditions (atmospheric pressure
Scenarios Depth of Duration of
and 20 °C). The values for the porosity and permeability were obtained
leakage leakage as described in Section 3.
source (m) (months) • Mechanical properties of the subsurface: particle density for each
0 No loss of containment (base case) 435 3 defined rock/soil type. The values for particle density were obtained
1 Leakage from cavern in Röt Claystone 395 3 from the website simetric.co.uk/si_materials.htm. Based on these
2 Scenario 1 with old permeable well in vicinity 395 3 values and the values for porosity, the STOMP model calculated the
3 Leakage from well below Hydrogeological base in 156 1
specific storativity, compressibility, and a tortuosity function for each
Muschelkalk Formation
4 Scenario 3 with old permeable well in vicinity 156 1 defined rock/soil type.
5 Leakage from well below hydrogeological base in 60 • Saturation properties of the saturation-capillary pressure function for
North Sea Supergroup each defined rock/soil type. For this function, the Van Genuchten
6 Scenario 5 with old permeable well in vicinity 60 water retention curve was used (van Genuchten, 1980). The Van
7 Leakage from well above hydrogeological base 20 1
Genuchten function was chosen because it assumes that the wetting
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 157

Table 3
Results for scenarios 0, 6 and 7 of the STOMP modelling, depth to top and bottom of NAPL (m), lateral spread of NAPL (m), maximum gas oil saturation in NAPL outside cavern (%), and
maximum gas oil saturation in phreatic groundwater (%).

Characteristics of gas oil NAPL Time horizons

1 day 1 week 1 month 3 months 1 year 5 year 30 year 150 year

Scenario 0
Depth to top NAPL (m) 436 436 436 436 436 436 436 436
Depth to bottom NAPL (m) 472 472 472 472 472 472 472 472
Spread NAPL (m) 105 105 105 105 105 105 105 105
Max. gas oil saturation in NAPL (%) 0.988 0.988 0.988 0.988 0.988 0.989 0.990 0.991
Max. saturation phreactic gw (%) 0 0 0 0 0 0 0 0

Scenario 6
Depth to top NAPL (m) 62 62 62 62 62 62 62 62
Depth to bottom NAPL (m) 64 64 64 64 64 64 64 64
Spread NAPL (m) 10 10 10 10 10 10 10 20
Max. gas oil saturation in NAPL (%) 0.002 0.015 0.036 0.036 0.036 0.036 0.037 0.038
Max. saturation phreactic gw (%) 0 0 0 0 0 0 0 0

Scenario 7
Depth to top NAPL (m) 20 18 14 8 0 0 0 0
Depth to bottom NAPL (m) 22 22 22 22 22 22 20 14
Spread NAPL (m) 10 10 30 30 30 50 70 90
Max. gas oil saturation in NAPL (%) 0.010 0.060 0.080 0.040 0.040 0.110 0.050 0.030
Max. saturation phreactic gw (%) 0.010 0.060 0.080 0.040 0.040 0.110 0.050 0.030

fluid drains from a porous medium whenever the capillary pressure 4.3. Results of multiphase modelling: effects of loss of containment
exceeds zero.
• The aqueous relative permeability for each defined rock/soil type. For The results of the STOMP model consist of the top and the bottom of
this purpose the Mualem function (Mualem, 1976) was chosen since the gas oil NAPL (metres below ground), the lateral spread of the NAPL
the aqueous relative permeability is dependent on the saturation (m), the maximum gas oil saturation in the NAPL (%), and the maximum
function (van Genuchten, 1980). saturation in the phreatic groundwater (%). From these values, which
• The NAPL relative permeability for each defined rock/soil type. Here were calculated for all 7 scenarios and the base case (no loss of contain-
too, the Mualem function was used because of its dependency on the ment), the effects of loss of containment could be determined. In the
Van Genuchten function. base case (scenario 0), there is no loss of containment, the model
• The physical properties of the gas oil: the gas oil to be injected was calculations show that during the 150-year modelling period the gas
assumed to have a density ranging from 820 to 860 kg m−3 (average oil saturation in the first metre of rock salt above the cavern is less
of 840 kg m−3) and a viscosity ranging from 1.64 × 10−3 Pa·s to than 1% and no gas oil was detected further from the cavern (see
3.87 × 10−3 Pa·s (average of 2.76 × 10−3 Pa·s). These values were Table 3). In the case of gas oil leakage from any point below the
obtained from Appendix B of the gas oil 10 ppm product specifications hydrogeological base, the model calculations show that during the
provided by the North Sea Group. 150-year modelling period the gas oil migrates only up to 25 m from
the point of leakage and does not contaminate the upper, phreatic
Table 1 shows all parameter values of the properties of each rock groundwater bodies in the Twente area. As an example, Fig. 5 and
type in the schematisation of the subsurface of the Twente area. Table 3 show the STOMP results for scenario 6 for the 150-year time
Boundary and initial conditions must also be stipulated before the point. Due to the low porosity and permeability of the geological layers,
model can be run. They are obtained by appropriately conceptualising upward migration is largely precluded and the gas oil does not reach the
the physical problem and translating that conceptualisation into bound- nearby structures that have greater porosity and permeability (fault or
ary and initial condition form. The initial conditions were defined for “old” well). However, leakage from the well above the hydrogeological
time step T = 0 and consisted of a hydrostatic pressure at the lower base in the phreatic aquifer does cause an immediate risk of contamina-
boundary of the model domain and a pressure gradient in the vertical tion of the upper groundwater bodies (scenario 7). After 1 year the gas
direction towards the upper boundary of the model domain. The pres- oil NAPL reaches the phreatic groundwater level and starts to spread out-
sure in the lowermost layer was calculated from the depth and the wards (reaching up to 90 m from the leakage point after 150 years).
water density (5.2 × 106 Pa). At the upper boundary the pressure is After 5 years the maximum saturation of gas oil in the pores of the sed-
identical to atmospheric pressure (1.01325 × 106 Pa) and therefore an iments is reached (0.11%), but because the NAPL continues to spread
unsaturated zone is created in the phreatic model layer. Another initial away from the point of leakage, the saturation within the NAPL reduces
condition is the NAPL pressure. In the lowest model layer, an NAPL pres- as it becomes partially trapped in the pores (see Table 3 and Figure 6).
sure was defined in order to obtain a layer completely saturated with
NAPL (5.2 × 106 Pa). In the rest of the model, the NAPL pressure was
defined as zero. The boundary conditions are constant for the whole 5. Semi-quantitative risk assessment
simulation period, and only at the upper boundary (the uppermost
layer) was the aqueous pressure defined to be equal to the atmospheric The first step in identifying and evaluating potential causes and
pressure (this is the hydrostatic pressure) and the NAPL was assigned a effects, and building scenarios is to elicit expert opinion (Wildenborg
zero flux. et al., 2005; Knol et al., 2010). The team of experts must be carefully
The seven flow scenarios were simulated with the STOMP model of composed, to minimise subjectivity and ensure confidence in their
the Twente area. Additionally, the situation without loss of containment judgement. The experts assign probabilities and impacts to potential
was modelled (scenario 0). Model results were generated for 10 points causes and effects, and resulting scenarios are semi-quantitatively.
in time after gas oil leakage occurs: 1 day, 1 week, 1 month, 3 months, Quantification of the impact of individual scenarios is possible during
1 year, 5 years, 30 years, 60 years, 100 year and 150 years. subsequent quantitative modelling.
158 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

Fig. 5. Visualisation of the STOMP results of scenario 6, effects of leakage after 150 years: leakage from well below hydrogeological base in North Sea Supergroup with old permeable well in
vicinity.

In risk assessment, risk is defined as Twente region (i.e. a working Containment Concept). Any deviations
from the base state such as the presence of leaking faults, unexpected
Risk ¼ Probability of Failure  Impact: presence of permeable layers or a too thin salt roof/pillar must be
defined as different base states. Events may occur that cause the base
In this study, the probability of failure was defined as the probability state to evolve into other states as defined by Table 4. Examples are
of gas oil leakage (the incident) and the impact was defined as the failure of mechanical equipment such as packer or casing, mechanical
harmful effects on health, safety and environment resulting from the failure of the cavern (formation of cracks, instability) or operational
contamination of groundwater. failure (high pressure).
The concept of “probability” we adhere to in this paper is the The Markov Chain ends with so-called “absorbing states” which
Bayesian one. It is rooted in our knowledge position, as opposed to the are leakage states. We distinguished three different leakage states
frequentist concept, which defines probability as an experimental based on the depth at which the leakage occurs; leakage above the
limit. The two interpretations are different, the mathematics are not. hydrogeological base, leakage from the hydrogeological base to just
The Bayesian concept has wider applicability. For instance, it allows above the cavern and leakage at cavern depth. These leakage states
one to speak of probabilities of hypotheses, which frequentist probability were separated according to the different impacts in the risk analysis;
cannot handle. The probabilities we used in the Markov Chain approach the impacts differ depending on the depth at which the leakage occurs.
(see below) are mostly based on expert judgement. That too requires Leakage above the hydrogeological boundary may lead directly to
the Bayesian stance on probability. For a thorough discussion on the groundwater contamination, while leakage at the well or at cavern
foregoing, see Jaynes (2003). depth will not immediately lead to contamination.
We limited ourselves to the underground risks involved (namely the
risk of contamination of groundwater). The surface risks such as leakage 5.1. Expert elicitation
of gas oil at the surface site, risk of explosion and risk to human health
were considered in the Environmental Impact Assessment (Tauw, As it is impossible to accurately know all circumstances in a well,
2013). The probability of gas oil leakage was calculated using two cavern and the subsurface and their development, we use probability
approaches. First, we used a Markov Chain approach (see Appendix A) theory for our formalisation. As phrased above, each situation or combi-
to form comprehensive scenarios based on occurring events or states. nation of situations can be considered a “state”, and some states have a
The resulting scenarios, in broad outline, were evaluated using input certain degree of likelihood that they will progress to certain other
from experts in the fields of geology, geohydrology, geomechanics and states. We defined the transition probabilities p(i,j) from state “i” to
well integrity. The impact was assessed by numerical modelling that state “j” for all states “i” and “j”. Having defined these transition proba-
simulated the flow of leaked gas oil (horizontal and vertical pathways) bilities, we could then compute how likely a certain state is at any time
into the groundwater (harmful effect). after the start date.
The starting point of Markov Chain analysis is a “base state” which The transition probabilities were based on expert knowledge. The
here is taken to be a properly functioning gas oil storage system in the experts were questioned about the likelihood of a transition from
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 159

Fig. 6. Visualisation of the STOMP results of scenario 7, effects of leakage after150 years: leakage from the well above the hydrogeological base. Lower figure is a close-up of the rectangle in
the upper figure, and shows the spread of the gas oil LNAPL on top of the phreatic groundwater level in more detail.

one state to another. They were asked to estimate this likelihood as the analysis. Fig. 7 shows the transition probabilities per year for the 27
probability of the transition occurring either in percentages or qualita- different states (Table 4) for the base case (i.e. a working Containment
tive statements as negligible, low, medium and high which were then Concept). The probability of a state onto itself (i.e. nothing changes)
translated into percentages (Figure 8) within a certain timeframe, is largest. The leakage states (# 25–27) are also clearly visible: for
expressed in years. These answers were converted to a likelihood example, the transition from an unstable cavern (state 6) to leakage at
expressed as percentage per year. Both the expert knowledge and the the well (state 26) and the cavern (state 27) is relatively high: 0.3 prob-
translation are qualitative. Hence, we had to investigate how the system ability per year.
evolves under a range of acceptable values for these transition probabil- The results of the Markov Chain analysis can be presented as proba-
ities. The results are general for the area considered and give insight into bilities of leakage for the three end absorbing states: leakage at the
what can be expected. All the transitions are expected to develop suffi- hydrogeological base, leakage around the well and leakage at cavern
ciently slowly, as we assume that the well is handled carefully and that depth. For the base case, the probability of leakage is negligible at all
the materials and operational methods used are state-of-the-art. For levels. The different malfunction states may occur at the cavern level
each specific storage case, we advise a case-specific quantitative risk or at the well. Since the base state does not include permeable faults
160 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

Table 4 large extent, be instrumental in preventing or lowering the risk of


Description of the 27 states considered in the Markov Chain analysis. leakage.
# state State description
5.2. Scenario assessment
1 Base case
2 Failed cementation
3 Failed packer The Markov Chain analysis gives insight into the effects of different
4 Failed casing states upon each other and the effects of human intervention and faulty
5 High pressure
site characterisation (presence of unknown faults, permeable layers
6 Unstable cavern
7 Formation of cracks etc.). From the result of the Markov Chain analysis eight scenarios
8 Failed casing & cementation were constructed from the most likely paths to leakage (highest transi-
9 Failed casing & cementation & packer tion probabilities):
10 High pressure & failed packer
11 High pressure & failed casing 1. Base case, working Containment Concept
12 Unstable cavern & failed packer 2. Well failure resulting in leakage below the hydrogeological base
13 High pressure & failed cementation 3. Well failure resulting in leakage above the hydrogeological base
14 High pressure & unstable cavern
15 High pressure& formation cracks
4. Unstable cavern resulting in cavern leakage
16 Unstable cavern & failed cementation 5. Unstable cavern resulting in well leakage
17 Unstable cavern & failed casing 6. High pressure resulting in leakage at cavern level
18 Unstable cavern & formation of cracks 7. High pressure resulting in well failure and leakage below the
19 Unstable cavern & failed packer & failed casing
hydrogeological base
20 Unstable cavern & failed cementation & failed casing & failed packer
21 High pressure & failed packer & unstable cavern 8. High pressure resulting in well failure and leakage above the
22 High pressure & failed packer & formation of cracks hydrogeological base.
23 High pressure & failed casing & unstable cavern
24 High pressure & failed casing & formation of cracks For each scenario control measures were defined, such as pressure
25 Leakage above hydrogeological boundary monitoring inside the cavern, monitoring of the annulus pressure,
26 Leakage around well monitoring gas oil and brine levels, sonar measurements (for observing
27 Leakage around cavern
the shape of the cavern) and safeguards to avert high pressure.
These scenarios and their control measures were presented to eight
experts/expert groups in the fields of geology, geomechanics and well
integrity. They were asked to fill the risk matrix according to their
expertise or leave the matrix empty if their expertise was not sufficient.
or layers that could transport the gas oil above the hydrogeological Fig. 8 shows the results of the scenario assessment in the form of a risk
level, the only pathway of leakage to the hydrogeological level is matrix. In this matrix the sub-classes for risk are defined as negligible,
the well itself. According to the experts, the probability of this occurring low, medium and high with associated probabilities and consequences
is low, hence we obtain a negligible probability of leakage at the of the leakage. The consequences of leakage (loss of containment) are
hydrogeological level. The existence of a permeable fault that intersects ranked according to the relative area (smaller or larger than 1 m3)
the cavern increases the probabilities of leakage at cavern depth and and timescale of the leakage (less than 1 month or 3 months) before
around the well to medium and high, respectively, over a period of the leakage is repaired. Since risk is defined as probability of failure ×
30 years. The existence of too thin a salt roof or too thin a salt pillar impact, the higher risks are associated with higher probability and larger
will lead to instability of the cavern and increased probabilities of areas and timescales of the leakage. Each coloured polygon indicates
leakage at cavern depth and around the well. This is mainly due to up- one expert or exert group; the number in the polygon indicates the
ward migration of the unstable cavern. Based on several different inputs scenario to which the risk refers. The arrow indicates the decrease in
for the transition probabilities we find that for a 30-year period the risk if control measures are taken (human intervention). Considering
probabilities for fluid release at shallow depths range from 0.04% to human intervention, scenarios 1, 2, 3, 5, 7 and 8 all show a negligible
0.3%, while for fluid release at cavern depth the probabilities range to low risk of leakage. For scenarios 4, 5 and 6 the loss of containment
from 0.07% to 2.5%. If we allow for worst-case estimates for the transi- will be larger and hence also the risk. Without human intervention,
tion probabilities, we obtain 2.6% probability of fluid release at shallow scenarios 4, 5 and 6 also yield the highest risk of leakage.
depths and 8.6% probability of fluid release at cavern depth. The proba- Overall, both the loss of containment and the probability of occur-
bility of fluid release at shallow depths is similar to that calculated by rence are assessed as higher if there is no human intervention, except
Veil et al. (1998) for the disposal of contaminated oilfield waste in salt for the base case, where the risk is considered negligible in both cases,
caverns (NORM study). The probability of fluid release at cavern depth assuming a working Containment Concept and monitoring system (no
is 3–5 times lower than the probabilities calculated by Veil et al. human error). There appears to be little consensus on the risks in the
(1998). But given that the NORM study was based on a specific storage case of no human intervention, indicating that it is more difficult for
site, whereas our study is general for the Twente area, the derived prob- experts to assess the risks of no human intervention. One expert opined
abilities of fluid release are not dissimilar. Since both studies are based that “no human intervention” was unrealistic and hence risk assess-
on expert knowledge, this provides a measure of confidence in our ment unnecessary. The risk assessment indicates that human interven-
translation of the expert knowledge into probabilities. tion is vital for minimising the risk. Therefore, in order to detect a loss of
So far, transition probabilities without human intervention have containment in an early stage, there must be continuous comparison of
been considered. This is unrealistic, however. In reality, there are well head pressures when the brine–hydrocarbon interface is in the
monitoring or mitigation tools which can and will be used to prevent cavern chimney, plus continuous monitoring of pressure at the bottom
leakage or act upon observed leakage. If the effects of human interven- of the well, pressure and flow in the well annulus, gas oil level, and brine
tion are taken into account, the probabilities of leakage are reduced and gas oil inflow and outflow. Additionally, the shape and extent of the
considerably. In other words, monitoring and mitigation are essential cavern should be monitored using sonar at intervals of at least 5 years,
in the prevention of leakage. This fits in with the historical cases of ideally before initial gas oil injection and after each gas oil extraction.
incidents (Evans, 2008; Evans and Chadwick, 2009): human error A change in the shape of the cavern will indicate a possibly instability
accounted for 94.4% of the problems which have been reported in Evan's of the cavern as well as possible additional salt dissolution. Both are
studies. In our risk assessment we see that human vigilance will, to a indications of leakage and require human intervention. Also, industry
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 161

Fig. 7. Transition probabilities for the base case: from state i (x-axis) to state j (y-axis) (definition of states in Table 4) for the period of one year. A 0.3 probability means a 30% probability in
one year. The scale of the colour bars is not the same in all four figures. The largest transition probabilities are given for the probability of state i onto itself as clearly shown in the top left
figure. In the bottom left figure, the transition probabilities to the leakage states (25–27) are clearly shown. In this figure too, there is a relatively high probability of transition from an
unstable cavern (state 6) to leakage at the well (state 26) and from the cavern (state 27).

standard mechanical integrity tests (MIT) of the day should be with gas oil injection and storage in salt caverns is low, an extensive
performed to monitor the integrity of the well during brine/gas oil monitoring plan should be formulated to ensure the long-term stability
production and injection. In addition, we recommend establishing a of the storage cavern and wells and containment of the gas oil. It should
baseline measurement of gas oil saturation and concentration of monitor the inflow and outflow at the wells, the pressures in the storage
dissolved gas oil components in groundwater prior to injection and system, the dimensions and shape of the caverns (regular sonar sur-
the monitoring of groundwater quality when/if gas oil leakage occurs, veys) and the integrity of the well to prevent leakage from the storage
as well as designing a contingency soil and groundwater remediation system through the causes identified in this paper and those based
plan. upon published incidents and accounts of release scenarios. Monitoring
This generic semi-quantitative risk assessment resulted in a check- should also be deployed to detect a loss of containment, especially
list of requirements (van Duijne et al., 2011) for the feasibility of gas groundwater contamination. Also, we advise preparing a contingency
oil storage in a specific salt cavern in the Twente area. This checklist soil and groundwater remediation plan. Finally, since human error is
was used for the cavern-specific risk analysis. the primary cause of incidents at salt storage facilities, it is essential to
have a robust management plan that includes enough safeguards to
6. Conclusions minimise the probability of human error.

A semi-quantitative assessment of the risks associated with storage Acknowledgements


of gas oil in an abandoned brine cavern in the eastern Netherlands
indicated that the risk of ground water contamination is low. This is This work was funded by Akzo Nobel BV, Hengelo, The Netherlands
due mainly to the low permeability of rocks in the area (rock salt, and carried out by Deltares and TNO. We thank Gerco Hoedeman, Laura
claystones). Consequently, there are a limited number of pathways for Wasch, Thijs Boxem and Wouter Meindertsma, TNO, who assisted in the
the stored product to migrate upward towards the groundwater level. project (specifically Figure 8, the checklist and revision). We also thank
Notwithstanding our conclusion that the risk of leakage associated Joy Burrough who advised on the English.
162 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

High

Breach of containment
5
6
>1m 2, 3 4
months

Medium
5 6 4
4
>1m 2, <1
4 6
month

Low
6 4 6 3
8
<1m 2, 3 2
5
7
months

Negligible
8 3 2 5
1
<1m 2, <1 7 2 3
1 1 5 7 8
month

Negligible Low Medium High

<0.001 0.001-0.01 0.01-0.1 >0.1

Probability

Fig. 8. Risk matrix for the eight scenarios. The background colours in the grid cells indicate risk: from negligible (blue) to low (green) to medium (orange) to high (red). The numbers in
polygons refer to the eight scenarios and the arrows show the change in risk from a situation in which there is no human intervention to one in which control measures are implemented.
The colours of the polygons indicate experts/expert groups: orange = experts who described the risk assuming there is always human intervention, claiming no human intervention is
unrealistic; blue and purple = two experts/expert groups assuming no human intervention and assuming human intervention as described by the arrows.

Appendix A. Markov Chains by a team of experts. It is obviously important that the relevant set
include “final” state(s) that describe the undesired situation(s). These
Given that a storage system needs to always operate safely it is states are considered absorbing: once reached there is no way back
necessary to ascertain what might cause the system to fail and what without drastic, wholesale mitigation.
are the risks. These risks have to be quantified. As the risk of an unwanted Given that states can evolve into other states, we make the practical
effect is defined in terms of the probability of an event and of this event and reasonable assumption that the likelihood of a transition per unit
having that effect, we should investigate both the probability of the event time depends solely on the state the system occupies. In that case we
and also the effect. This appendix deals with event probability. can apply the machinery of Absorbing Markov Chains to arrive at a quan-
We begin by noting that a system can be in various “states”. There is titative statement for the occupation probabilities involved. To that end
a certain probability that these states will evolve into other states. As a the investigator has to define transition probabilities between the states
first step, we need to define the “base state” from which the other states as he or she defined them. The unit of time is typically one year. Infor-
may ensue. The base state in this paper is a working containment situa- mation from the experts will enable the investigator to derive typical
tion as described in the main text. values for these transition probabilities. How expert elicitation should
Any deviations from the Containment Concept, such as the unfore- be done will not concern us in this appendix. Suffice it to say that it is
seen presence of leaky faults, permeable layers, a too thin salt roof or a not necessarily a very straightforward process, as is known from the
too thin salt pillar will have to be defined as different possible states psychological literature on the subject as quoted in Nepveu et al. (2009).
for the analysis. But also any conceivable combination is a possible Ultimately, then, one obtains a transition matrix A(i, j) containing
state. So, if there are N “monolithic” states there are 2N-1 combinations the transition probabilities from the states i = Nj. A is a square matrix
possible, as simple combinatorics shows. Not all of these might be of dimension S × S where S is the number of states deemed relevant.
deemed relevant, and the investigator has the task to trim all possible The final states as mentioned above are absorbing; once they are
combinations to those that are potentially relevant. The investigator reached they are never left by the system. All the other states are called
does so taking account of a previous qualitative risk analysis performed transient states.
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 163

Especially interesting for elicitation purposes are diagonal elements; Table B.1 (continued)
they are related to the retention time (=uninterrupted occupation From # state To # state Base case Var1 Var2 Var3 Var4
time). The expert can certainly produce a rough number or a range for 1 7 0.00030 0.00030 0.00030 0.00030 0.00030
this retention time. This can be used to compute the transition probabil- 2 8 0.00200 0.00200 0.00200 0.00200 0.00200
ity A(i,i) as 2 13 0.00010 0.00010 0.00010 0.00010 0.00010
2 16 0.00010 0.00010 0.00010 0.00010 0.00010
3 10 0.01000 0.00300 0.00300 0.00300 0.01000
3 12 0.00300 0.00030 0.00030 0.00030 0.00300
−1
Aði; iÞ ¼ 1:−Tretention : 4 8 0.00500 0.00300 0.00300 0.00300 0.00500
4 11 0.00500 0.00300 0.00300 0.00300 0.00500
4 17 0.00050 0.00030 0.00030 0.00030 0.00050
5 10 0.00500 0.00500 0.00500 0.00500 0.00500
5 11 0.00300 0.00300 0.00300 0.00300 0.00300
So, if the diagonal element is zero it means that after one year the
5 13 0.00050 0.00050 0.00050 0.00050 0.00050
state i is definitely abandoned if initially occupied; If the diagonal 5 14 0.00030 0.00030 0.00030 0.00030 0.00030
element equals unity the state is absorbing: it will never be left, when 5 15 0.10000 0.10000 0.10000 0.10000 0.10000
occupied. This last condition will hold true for the “final” states. 6 12 0.10000 0.10000 0.10000 0.10000 0.02000
6 14 0.00200 0.00200 0.00200 0.00200 0.00200
Matrix A should have the formal properties of each transition matrix
6 16 0.00500 0.00500 0.00500 0.00500 0.00500
in that 6 17 0.03000 0.03000 0.03000 0.03000 0.02000
6 18 0.10000 0.10000 0.10000 0.10000 0.02000
6 19 0.00100 0.00100 0.00100 0.00100 0.00100
X 6 20 0.00050 0.00050 0.00050 0.00050 0.00050
A j ðiÞ≥ 0 for all i; j and j
A j ðiÞ ¼ 1 for all i: ðA:1Þ
6 26 0.35000 0.35000 0.35000 0.35000 0.10000
6 27 0.35000 0.35000 0.35000 0.35000 0.10000
7 15 0.00030 0.00030 0.00030 0.00030 0.00030
7 18 0.00030 0.00030 0.00030 0.00030 0.00030
The second property simply expresses the fact that the system must 7 25 0.00300 0.00300 0.00300 0.00300 0.00300
7 26 0.03000 0.03000 0.03000 0.03000 0.03000
necessarily be in one of the states defined — there are no states other
7 27 0.03000 0.03000 0.03000 0.03000 0.03000
than the ones we defined. When the initial state vector is v(0) = 8 20 0.00030 0.00030 0.00030 0.00030 0.00030
(1, 0, 0, ….) – this probability vector represents the fact that the initial 8 25 0.10000 0.10000 0.10000 0.10000 0.10000
state is the base case with probability one – then the state vector after 8 26 0.10000 0.10000 0.10000 0.10000 0.10000
n time units is equal to 8 27 0.10000 0.10000 0.10000 0.10000 0.10000
9 20 0.00030 0.00030 0.00030 0.00030 0.00030
9 25 0.10000 0.10000 0.10000 0.10000 0.10000
9 26 0.10000 0.10000 0.10000 0.10000 0.10000
n
vðnÞ ¼ vð0ÞA ðA:2Þ 9 27 0.10000 0.10000 0.10000 0.10000 0.10000
10 21 0.00030 0.00030 0.00030 0.00030 0.00030
10 22 0.00030 0.00030 0.00030 0.00030 0.00030
10 26 0.04000 0.03000 0.03000 0.03000 0.04000
and describes the occupation probability for the different states under
10 27 0.04000 0.03000 0.03000 0.03000 0.04000
scrutiny. Since the transition probabilities obtained from the experts 11 23 0.00030 0.00030 0.00030 0.00030 0.00030
cannot be supposed to have “rock solid” status, some experimentation 11 24 0.00030 0.00030 0.00030 0.00030 0.00030
is needed, as a kind of sensitivity analysis. This was done in the work 11 26 0.03000 0.03000 0.03000 0.03000 0.03000
this paper describes. 11 27 0.03000 0.03000 0.03000 0.03000 0.03000
12 19 0.05000 0.05000 0.05000 0.05000 0.05000
Note that in an absorbing Markov Chain the final states are reached
12 21 0.00300 0.00300 0.00300 0.00300 0.00300
eventually. More precisely: lim An exists as n goes to infinity and in this 12 26 0.10000 0.10000 0.10000 0.10000 0.10000
limit all transient states are evacuated and at least some if not all of the 12 27 0.10000 0.10000 0.10000 0.10000 0.10000
final states are occupied with probabilities such that the total sum of 13 26 0.10000 0.10000 0.10000 0.10000 0.10000
13 27 0.10000 0.10000 0.10000 0.10000 0.10000
these equals unity. The probabilities of these states' occupation depend
14 21 0.01000 0.01000 0.01000 0.01000 0.01000
on the structure of A. The non-zero probabilities of reaching the final 14 23 0.01000 0.01000 0.01000 0.01000 0.01000
state(s) induce them to be occupied eventually. This makes perfect 14 26 0.10000 0.10000 0.10000 0.10000 0.10000
sense mathematically. It does not prove that in practice all must go 14 27 0.10000 0.10000 0.10000 0.10000 0.10000
wrong. It merely shows what the system has in store for us if we do 15 22 0.03000 0.03000 0.03000 0.03000 0.03000
15 24 0.01000 0.01000 0.01000 0.01000 0.01000
not interfere at critical moments. Timescales on which that can be sen-
15 27 0.05000 0.05000 0.05000 0.05000 0.05000
sibly done follow from the analysis as well (Nepveu et al., 2009). 16 20 0.10139 0.00300 0.03000 0.10000 0.10000
17 19 0.25000 0.50000 0.50000 0.50000 0.25000
Appendix B. Input and output of absorbing Markov Chain analysis 17 20 0.03000 0.03000 0.03000 0.03000 0.03000
17 23 0.00300 0.00300 0.00300 0.00300 0.00300
17 27 0.03000 0.03000 0.03000 0.03000 0.03000
Table B.1 to Table B.4 show some values for the transition probabil- 18 27 0.50000 0.50000 0.50000 0.50000 0.50000
ities of different scenarios. 19 20 0.03000 0.03000 0.03000 0.03000 0.03000
19 26 0.15000 0.15000 0.15000 0.15000 0.15000
19 27 0.30000 0.30000 0.30000 0.30000 0.30000
20 26 0.15000 0.15000 0.15000 0.15000 0.15000
Table B.1 20 27 0.30000 0.30000 0.30000 0.30000 0.30000
Transition probabilities for the base case and four variants of the base case. 21 26 0.15000 0.15000 0.15000 0.15000 0.15000
21 27 0.30000 0.30000 0.30000 0.30000 0.30000
From # state To # state Base case Var1 Var2 Var3 Var4
22 26 0.05000 0.05000 0.05000 0.05000 0.05000
1 2 0.00200 0.00200 0.00200 0.00200 0.00200 22 27 0.10000 0.10000 0.10000 0.10000 0.10000
1 3 0.00400 0.00400 0.00400 0.00400 0.00400 23 26 0.15000 0.15000 0.15000 0.15000 0.15000
1 4 0.00200 0.00200 0.00200 0.00200 0.00200 23 27 0.30000 0.30000 0.30000 0.30000 0.30000
1 5 0.00100 0.00100 0.00100 0.00100 0.00100 24 26 0.05000 0.05000 0.05000 0.05000 0.05000
1 6 0.00030 0.00030 0.00030 0.00030 0.00030 24 27 0.10000 0.10000 0.10000 0.10000 0.10000
164 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

Table B.2 Table B.2 (continued)


Transition probabilities for 4 worst-case scenarios. From To Worst-case 1 Worst-case 2 Worst-case 3 Worst-case 4
# state # state
From To Worst-case 1 Worst-case 2 Worst-case 3 Worst-case 4
# state # state 21 27 0.3 0.3 0.3 0.3
1 2 0.004 0.004 0.004 0.004 22 26 0.05 0.05 0.05 0.05
1 3 0.006 0.006 0.006 0.006 22 27 0.1 0.1 0.1 0.1
1 4 0.004 0.004 0.004 0.004 23 26 0.15 0.15 0.15 0.15
1 5 0.002 0.002 0.002 0.002 23 27 0.3 0.3 0.3 0.3
1 6 0.0003 0.003 0.00003 0.00003 24 26 0.05 0.05 0.05 0.05
24 27 0.1 0.1 0.1 0.1
1 7 0.003 0.0003 0.00003 0.00003
2 8 0.04 0.04 0.04 0.004
2 13 0.0001 0.0001 0.0001 0.0001
2 16 0.0001 0.0001 0.0001 0.0001
3 10 0.01 0.01 0.01 0.01
3 12 0.003 0.003 0.003 0.003
4 8 0.05 0.05 0.05 0.005
4 11 0.005 0.005 0.005 0.005
4 17 0.0005 0.0005 0.0005 0.0005 Table B.3
5 10 0.005 0.005 0.005 0.005 Transition probabilities for 4 cases: permeable fault, permeable layer, thin salt pillar and
5 11 0.003 0.003 0.003 0.003 thin salt roof.
5 13 0.0005 0.0005 0.0005 0.0005
From # state To # state Permeable fault Permeable layer Thin pillar Thin roof
5 14 0.0003 0.0003 0.0003 0.0003
5 15 0.1 0.1 0.1 0.1 1 2 0.002 0.002 0.002 0.002
6 12 0.1 0.1 0.1 0.1 1 3 0.004 0.004 0.004 0.004
6 14 0.002 0.002 0.002 0.002 1 4 0.002 0.002 0.002 0.002
6 16 0.005 0.005 0.005 0.005 1 5 0.001 0.001 0.001 0.001
6 17 0.03 0.03 0.03 0.03 1 6 0.0003 0.0003 0.5 0.5
6 18 0.1 0.1 0.1 0.1 1 7 0.0003 0.0003 0.0003 0.0003
6 19 0.001 0.001 0.001 0.001 2 8 0.002 0.002 0.002 0.002
6 20 0.0005 0.0005 0.0005 0.0005 2 13 0.0001 0.0001 0.0001 0.0001
6 26 0.35 0.35 0.35 0.35 2 16 0.0001 0.0001 0.0001 0.0001
6 27 0.35 0.35 0.35 0.35 3 10 0.01 0.01 0.01 0.01
7 15 0.0003 0.0003 0.0003 0.0003 3 12 0.003 0.003 0.003 0.003
7 18 0.0003 0.0003 0.0003 0.0003 4 8 0.005 0.005 0.005 0.005
7 25 0.003 0.003 0.003 0.003 4 11 0.005 0.005 0.005 0.005
7 26 0.03 0.03 0.03 0.03 4 17 0.0005 0.0005 0.0005 0.0005
7 27 0.03 0.03 0.03 0.03 5 10 0.005 0.005 0.005 0.005
8 20 0.0003 0.0003 0.0003 0.0003 5 11 0.003 0.003 0.003 0.003
8 25 0.1 0.1 0.1 0.1 5 13 0.0005 0.0005 0.0005 0.0005
8 26 0.1 0.1 0.1 0.1 5 14 0.0003 0.0003 0.0003 0.0003
8 27 0.1 0.1 0.1 0.1 5 15 0.1 0.1 0.1 0.1
9 20 0.0003 0.0003 0.0003 0.0003 6 12 0.1 0.1 0.1 0.1
9 25 0.1 0.1 0.1 0.1 6 14 0.002 0.002 0.002 0.002
9 26 0.1 0.1 0.1 0.1 6 16 0.005 0.005 0.005 0.005
9 27 0.1 0.1 0.1 0.1 6 17 0.03 0.03 0.03 0.03
10 21 0.0003 0.0003 0.0003 0.0003 6 18 0.1 0.1 0.1 0.1
10 22 0.0003 0.0003 0.0003 0.0003 6 19 0.001 0.001 0.001 0.001
10 26 0.04 0.04 0.04 0.04 6 20 0.0005 0.0005 0.0005 0.0005
10 27 0.04 0.04 0.04 0.04 6 26 0.1 0.2 0.35 0.35
11 23 0.0003 0.0003 0.0003 0.0003 6 27 0.1 0.4 0.35 0.35
11 24 0.0003 0.0003 0.0003 0.0003 7 15 0.0003 0.0003 0.0003 0.0003
11 26 0.03 0.03 0.03 0.03 7 18 0.0003 0.0003 0.0003 0.0003
11 27 0.03 0.03 0.03 0.03 7 25 0.03 0.003 0.003 0.003
12 19 0.05 0.05 0.05 0.05 7 26 0.03 0.03 0.03 0.03
12 21 0.003 0.003 0.003 0.003 7 27 0.03 0.3 0.03 0.03
12 26 0.1 0.1 0.1 0.1 8 20 0.0003 0.0003 0.0003 0.0003
12 27 0.1 0.1 0.1 0.1 8 25 0.1 0.1 0.1 0.1
13 26 0.1 0.1 0.1 0.1 8 26 0.1 0.1 0.1 0.1
13 27 0.1 0.1 0.1 0.1 8 27 0.1 0.1 0.1 0.1
14 21 0.01 0.01 0.01 0.01 9 20 0.0003 0.0003 0.0003 0.0003
14 23 0.01 0.01 0.01 0.01 9 25 0.1 0.1 0.1 0.1
14 26 0.1 0.1 0.1 0.1 9 26 0.1 0.1 0.1 0.1
14 27 0.1 0.1 0.1 0.1 9 27 0.1 0.1 0.1 0.1
15 22 0.03 0.03 0.03 0.03 10 21 0.0003 0.0003 0.0003 0.0003
15 24 0.01 0.01 0.01 0.01 10 22 0.0003 0.0003 0.0003 0.0003
15 27 0.05 0.05 0.05 0.05 10 26 0.04 0.04 0.04 0.04
16 20 0.1 0.1 0.1 0.1 10 27 0.04 0.04 0.04 0.04
17 19 0.25 0.25 0.25 0.25 11 23 0.0003 0.0003 0.0003 0.0003
17 20 0.03 0.03 0.03 0.03 11 24 0.0003 0.0003 0.0003 0.0003
17 23 0.003 0.003 0.003 0.003 11 26 0.03 0.03 0.03 0.03
17 27 0.03 0.03 0.03 0.03 11 27 0.03 0.03 0.03 0.03
18 27 0.5 0.5 0.5 0.5 12 19 0.05 0.05 0.05 0.05
19 20 0.03 0.03 0.03 0.03 12 21 0.003 0.003 0.003 0.003
19 26 0.15 0.15 0.15 0.15 12 26 0.1 0.1 0.1 0.1
19 27 0.3 0.3 0.3 0.3 12 27 0.1 0.3 0.1 0.1
20 26 0.15 0.15 0.15 0.15 13 26 0.1 0.1 0.1 0.1
20 27 0.3 0.3 0.3 0.3 13 27 0.1 0.3 0.1 0.1
21 26 0.15 0.15 0.15 0.15 14 21 0.01 0.01 0.01 0.01
K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166 165

Table B.3 (continued) Table B.4 (continued)


From # state To # state Permeable fault Permeable layer Thin pillar Thin roof From To Monitor Monitor Monitor well Safe Some human
# state # state pressure pressure scenario action
14 23 0.01 0.01 0.01 0.01
& well
14 26 0.1 0.1 0.1 0.1
14 27 0.1 0.3 0.1 0.1 4 17 0.0003 0.0003 0.002 0.002 0.002
15 22 0.03 0.03 0.03 0.03 5 10 0.000001 0.000001 0.0001 0.0001 0.0001
15 24 0.01 0.01 0.01 0.01 5 11 0.000001 0.000001 0.0001 0.0001 0.0001
15 27 0.05 0.3 0.05 0.05 5 13 0.000001 0.000001 0.003 0.003 0.003
16 20 0.10139 0.10139 0.10139 0.10139 5 14 0.000001 0.000001 0.0003 0.0003 0.0003
17 19 0.25 0.25 0.25 0.25 5 15 0.000001 0.000001 0.003 0.003 0.003
17 20 0.03 0.03 0.03 0.03 6 12 0.1 0.1 0.003 0.003 0.003
17 23 0.003 0.003 0.003 0.003 6 14 0.002 0.002 0.0003 0.0003 0.0003
17 27 0.03 0.3 0.03 0.03 6 16 0.005 0.005 0.005 0.005 0.005
18 27 0.1 0.3 0.5 0.5 6 17 0.03 0.03 0.003 0.003 0.003
19 20 0.03 0.03 0.03 0.03 6 18 0.1 0.1 0.0005 0.0005 0.0005
19 26 0.2 0.15 0.15 0.15 6 19 0.001 0.001 0.0003 0.0003 0.0003
19 27 0.2 0.3 0.3 0.3 6 20 0.0005 0.0005 0.1 0.1 0.1
20 26 0.2 0.15 0.15 0.15 6 26 0.35 0.35 0.1 0.1 0.1
20 27 0.2 0.3 0.3 0.3 6 27 0.35 0.35 0.002 0.002 0.002
21 26 0.2 0.15 0.15 0.15 7 15 0.0003 0.0003 0.005 0.005 0.005
21 27 0.2 0.3 0.3 0.3 7 18 0.0003 0.0003 0.03 0.03 0.03
22 26 0.1 0.05 0.05 0.05 7 25 0.003 0.003 0.1 0.1 0.1
22 27 0.1 0.3 0.1 0.1 7 26 0.03 0.03 0.001 0.001 0.001
23 26 0.2 0.15 0.15 0.15 7 27 0.03 0.03 0.0005 0.0005 0.0005
23 27 0.2 0.3 0.3 0.3 8 1 0.000001 0.2 0.35 0.35 0.35
24 26 0.1 0.05 0.05 0.05 8 20 0.0003 0.0003 0.35 0.35 0.35
24 27 0.1 0.3 0.1 0.1 8 25 0.1 0.1 0.0003 0.0003 0.0003
1 25 0.01 0.000001 0.000001 0.000001 8 26 0.1 0.1 0.0003 0.0003 0.0003
1 26 0.05 0.000001 0.000001 0.000001 8 27 0.1 0.1 0.003 0.003 0.003
1 27 0.1 0.1 0.1 0.000001 9 1 0.000001 0.2 0.03 0.03 0.03
6 25 0.1 0.000001 0.000001 0.000001 9 20 0.0003 0.0003 0.03 0.03 0.03
10 25 0.04 0.000001 0.000001 0.000001 9 25 0.1 0.1 0.0003 0.0003 0.0003
11 25 0.03 0.000001 0.000001 0.000001 9 26 0.1 0.1 0.1 0.1 0.1
12 25 0.1 0.000001 0.000001 0.000001 9 27 0.1 0.1 0.1 0.1 0.1
13 25 0.1 0.000001 0.000001 0.000001 10 21 0.000001 0.000001 0.1 0.1 0.1
14 25 0.1 0.000001 0.000001 0.000001 10 22 0.000001 0.000001 0.0003 0.0003 0.0003
15 25 0.05 0.000001 0.000001 0.000001 10 26 0.000001 0.000001 0.1 0.1 0.1
15 26 0.05 0.000001 0.000001 0.000001 10 27 0.000001 0.000001 0.1 0.1 0.1
17 25 0.03 0.000001 0.000001 0.000001 11 23 0.000001 0.000001 0.1 0.1 0.1
17 26 0.03 0.000001 0.000001 0.000001 11 24 0.000001 0.000001 0.0003 0.0003 0.0003
18 25 0.1 0.000001 0.000001 0.000001 11 26 0.000001 0.000001 0.0003 0.0003 0.0003
18 26 0.1 0.000001 0.000001 0.000001 11 27 0.000001 0.000001 0.03 0.03 0.03
19 25 0.2 0.000001 0.000001 0.000001 12 6 0.000001 0.2 0.03 0.03 0.03
20 25 0.2 0.000001 0.000001 0.000001 12 19 0.05 0.05 0.0003 0.0003 0.0003
21 25 0.2 0.000001 0.000001 0.000001 12 21 0.003 0.003 0.0003 0.0003 0.0003
22 25 0.1 0.000001 0.000001 0.000001 12 26 0.1 0.1 0.03 0.03 0.03
23 25 0.2 0.000001 0.000001 0.000001 12 27 0.1 0.1 0.03 0.03 0.03
24 25 0.1 0.000001 0.000001 0.000001 13 26 0.000001 0.000001 0.05 0.05 0.05
13 27 0.000001 0.000001 0.003 0.003 0.003
14 21 0.000001 0.000001 0.1 0.1 0.1
14 23 0.000001 0.000001 0.1 0.1 0.1
14 26 0.000001 0.000001 0.1 0.1 0.1
14 27 0.000001 0.000001 0.1 0.1 0.1
15 22 0.000001 0.000001 0.01 0.01 0.01
15 24 0.000001 0.000001 0.01 0.01 0.01
15 27 0.000001 0.000001 0.1 0.1 0.1
16 6 0.000001 0.2 0.1 0.1 0.1
Table B.4 16 20 0.003 0.003 0.03 0.03 0.03
Transition probabilities for the human intervention scenarios. 17 6 0.000001 0.2 0.01 0.01 0.01
17 19 0.5 0.5 0.05 0.05 0.05
From To Monitor Monitor Monitor well Safe Some human
17 20 0.03 0.03 0.003 0.003 0.003
# state # state pressure pressure scenario action
17 23 0.003 0.003 0.5 0.5 0.5
& well
17 27 0.03 0.03 0.03 0.03 0.03
1 2 0.002 0.002 0.002 0.0002 0.003 18 27 0.5 0.5 0.003 0.003 0.003
1 3 0.004 0.004 0.004 0.0004 0.007 19 6 0.000001 0.2 0.03 0.03 0.03
1 4 0.002 0.002 0.002 0.0002 0.003 19 20 0.03 0.03 0.5 0.5 0.5
1 5 0.001 0.001 0.001 0.0001 0.002 19 26 0.15 0.15 0.03 0.03 0.03
1 6 0.0003 0.0003 0.0003 0.00003 0.0003 19 27 0.3 0.3 0.15 0.15 0.15
1 7 0.0003 0.0003 0.0003 0.00003 0.0003 20 6 0.000001 0.2 0.3 0.3 0.3
2 1 0.000001 0.2 0.2 0.2 0.05 20 26 0.15 0.15 0.15 0.15 0.15
2 8 0.002 0.002 0.2 0.2 0.05 20 27 0.3 0.3 0.3 0.3 0.3
2 13 0.0001 0.0001 0.2 0.2 0.05 21 26 0.000001 0.000001 0.15 0.15 0.15
2 16 0.0001 0.0001 0.2 0.2 0.000001 21 27 0.000001 0.000001 0.3 0.3 0.3
3 1 0.000001 0.2 0.2 0.2 0.000001 22 26 0.000001 0.000001 0.05 0.05 0.05
3 10 0.003 0.003 0.2 0.2 0.000001 22 27 0.000001 0.000001 0.1 0.1 0.1
3 12 0.0003 0.0003 0.2 0.2 0.000001 23 26 0.000001 0.000001 0.15 0.15 0.15
4 1 0.000001 0.2 0.2 0.2 0.000001 23 27 0.000001 0.000001 0.3 0.3 0.3
4 8 0.003 0.003 0.2 0.2 0.000001 24 26 0.000001 0.000001 0.05 0.05 0.05
4 11 0.003 0.003 0.2 0.2 0.000001 24 27 0.000001 0.000001 0.1 0.1 0.1
166 K. van Thienen-Visser et al. / Engineering Geology 168 (2014) 149–166

References Kunstman, A., Urbańczyk, K., 2008. Designing of the storage caverns for liquid products,
anticipating its size and shape changes during withdrawal operations with use of
Anderson, E.I., Bakker, M., 2008. Groundwater flow through anisotropic fault zones in unsaturated brine. SMRI 2008 Spring meeting, Technical Conference, 28–29 April,
multi-aquifer systems. Water Resour. Res. 44, W11433. http://dx.doi.org/10.1029/ Porto, Portugal.
2008WR006925. Kutchko, B.G., Strazisar, B.R., Dzombak, D.A., Lowry, G.V., Thaulow, N., 2007. Degradation
Bear, J., 1972. Dynamics of Fluids in Porous Media. Dover0-486-65675-6. of well cement by CO2 under geological sequestration conditions. Environ. Sci.
Bense, V.F., Person, M.A., 2006. Faults as conduit–barrier systems to fluid flow in Technol. 2007 (41), 4787–4792.
siliciclastic sedimentary aquifers. Water Resour. Res. 42, W05421. http://dx.doi.org/ Lenhard, R.J., Parker, J.C., 1987. Measurement and prediction of saturation–pressure
10.1029/ 2005WR004480. relationships in three-phase porous media. J. Contam. Hydrol. 1, 407–424.
Bérest, P., Brouard, B., 2003. Safety of salt caverns used for underground gas storage. Oil Lenhard, R.J., Oostrom, M., White, M.D., 1995. Modeling fluid flow and transport in
Gas Sci. Technol. 58, 361–384. variably saturated porous media with the STOMP simulator. 2. Verification and
Bérest, P., Brouard, B., Durup, J.G., 2001. Tightness tests in salt cavern wells. Oil Gas Sci. validation exercises. Adv. Water Resour. 18 (6), 365–373.
Technol. 56, 451–469. Magri, F., Akar, T., Gemici, U., Pekdeger, A., 2010. Deep geothermal groundwater flow in
BGR, 1998. Rock-mechanical expertise concerning the stability of the existing caverns and the Seferihisar–Balc¸ova area, Turkey: results from transient numerical simulations
for dimensioning of new caverns in the Hengelo field. Report (English Translation) of coupled fluid flow and heat transport processes. Geofluids 10, 388–405. http://
prepared by Bundesanstalt Fur Geowissenschaften und Rohstoffe (BGR), Hanover, dx.doi.org/10.1111/j.1468-8123.2009.00267.x.
Germany, for Akzo Nobel Salt, Hengelo, The Netherlands. 1998. Marsman, A., 2002. The Influence of Water Pollution on the Flow of Light Non Aqueous
BGR, 2004a. Rock-mechanical calculations for the stability and integrity of inline pillars Phase Liquids in Soil. (Phd Thesis) Wageningen University90-5808-734-4.
within parallel cavern rows. Report prepared by Bundesanstalt Fur Geowissenschaften Miyazaki, B., 2009. Well integrity: an overlooked source of risk and liability for under-
und Rohstoffe (BGR), Hannover, Duitsland for Akzo Nobel Salt B.V., Hengelo, The ground natural gas storage. Lessons learned from incidents in the USA. Geol. Soc.
Netherlands, November 2004. Lond. Spec. Publ. 313, 163–172. http://dx.doi.org/10.1144/SP313.11.
BGR, 2004b. Investigation in the bearing load of pillars in the Hengelo terrain of Akzo Mualem, Y., 1976. A new model for predicting the hydraulic conductivity of unsaturated
Nobel. Report prepared by Bundesanstalt Fur Geowissenschaften und Rohstoffe porous media. Water Resour. Res. 12, 513–522. http://dx.doi.org/10.1029/
(BGR), Hannover, Duitsland for Akzo Nobel Salt B.V., Hengelo, The Netherlands, WR012i003p00513.
March 2004. MWH, B.V., 2010. Salt mining possibilities in areas adjacent to the Hengelo brine field.
Boukhelifa, L., Moroni, N., James, S.G., Le Roy-Delage, S., Thiercelin, M.J., Lemaire, G., 2004. Prepared for Akzo Nobel Industrial Chemicals B.V., Feb 2010.
Evaluation of cement systems for oil and gas well zonal isolation in a full-scale Nepveu, M., Yavuz, F., David, P., 2009. FEP analysis and Markov chains. Energy Procedia I
annular geometry. SPE Paper 87195, IADC/SPE Drilling Conference Dallas, TX, U.S.A., 2009, 2519–2523.
March 2–4. Nieland, J.D., Ratigan, J.L., 2006. Geomechanical evaluation of two gulf coast natural as
Bouw, L., Oude Essink, G.H.P., 2003. Fluid flow in the northern Broad Fourteens Basin storage caverns. SMRI 2006 Spring meeting, 30 April–3 May, 2006, Brussels, Begium.
during Late Cretaceous inversion. Neth. J. Geosci.Geol. Mijnb. 82 (1), 55–69. NITG-TNO, 1998. Geologische atlas van de diepe ondergrond van Nederland, kaartblad X:
De Louw, 2006. Wateratlas Twente, De grond en oppervlaktewatersystemen van Regge Almelo – Winterswijk.
en Dinkel. Waterschap Regge en Dinkel/TNO Bouw en Ondergrond. Pöppelreiter, M., Borkhataria, R., Aigner, T., Pipping, K., 2005. Production from
Doe, T.W., Osnes, J.D., 2006. In situ stress and permeability tests in the Hutchinson salt Muschelkalk carbonates (Triassic, NE Netherlands): unique play or overlooked
and the overlying shale, Kansas. Technical Conference paper, Fall 2006 Conference, opportunity? Geological Society, London, Petroleum Geology Conference series, 6,
Rapid City, South Dakota, USA. pp. 299–315. http://dx.doi.org/10.1144/0060299.
Domenico, A.D., Schwartz, F.W., 1998. Physical and Chemical Hydrology. John Wiley and Querio, C.W., 1980. Design and construction of solution-mined caverns for LPG storage.
Sons, Inc.0-471-59762-7. Presented at the Fall Meeting of the Solution Mining Research Institute, Minneapolis,
Dufour, F.C., 1998. Grondwater in Nederland, Onzichtbaar water waarop wij lopen. NITG- Minnesota, October.
TNO90-6743-536-8. Saar, M.O., 2010. Review: geothermal heat as a tracer of large-scale groundwater flow and
Evans, D.J., 2008. Accidents at UFS sites and risk relative to other areas of the energy as a means to determine permeability fields. Hydrogeol. J. 19, 31–52. http://dx.doi.org/
supply chain, with particular reference to salt cavern storage. SMRI Fall 2008 Technical 10.1007/s10040-010-0657-2.
Conference, 13–14 October Galveston, Texas, USA. Tauw, 2013. Environmental Impact Assessment for the Storage of Gasoil in Existing Salt
Evans, D.J., Chadwick, R.A., 2009. Underground gas storage : worldwide experiences and Caverns, by Tauw B.V. Under Contract of AkzoNobel Industrial Chemicals B.V.
future development in the UK and Europe. Geological Society Special, Publication, Thoms, R.L., Gehle, R.M., 2000. A brief history of salt cavern use (keynote paper). In:
313. Geological Society of London, London, UK (283 pp.). Geertman, R.M. (Ed.), Proceedings of 8th World Salt Symposium, part 1. Elsevier
Eyermann, T.J., 1994. Status of the strategic petroleum reserves. SMRI 1994 Spring B.V., pp. 207–214.
meeting, Technical conference, 25–26 April, Houston, Texas. Van Duijne, H., Hendriks, D., van Thienen-Visser, K., Wildenborg, T., 2011. Report of
Folch, A., Mas-pla, J., 2008. Hydrogeological interactions between fault zones and alluvial Interviews and Workshop 2U-CC Project (Confidential) (Deltares Memo 1203390-
aquifers in regional flow systems. Hydrogeol. Process. 22, 3476–3487. http://dx.doi.org/ 000-BGS-0016).
10.1002/hyp.6956. Van Genuchten, M.T., 1980. A closed form equation for predicting the hydraulic conduc-
Gasda, S.E., Bachu, S., Celia, M.A., 2004. The potential for CO2 leakage from storage sites in tivity of unsaturated soils. Soil Sci. Soc. Am. J. 44, 892–898.
geological media: analysis of well distribution in mature sedimentary basins. Environ. Van Sambeek, L.L., Bérest, P., Brouard, B., 2005. Improvements in mechanical integrity
Geol. 46 (6–7), 707–720. tests for solution-mined caverns used for mineral production or liquid-product
Geowulf, 2008. Geological framework TWR area compliation report. Prepared for Akzo storage. Prepared for SMRI in Cooperation with MIT, May.
Nobel Industrial Chemicals B.V., 2008, vol 1 (report no: GL08.502). Veil, J., Smith, K., Tomasko, D. Elcock, Williams, G., Blunt, D., 1998. Disposal of NORM-
GeoWulf, 2010. Geological analysis of the Marssteden area. Twente Region for AkzoNobel contaminated oil field wastes in salt caverns — legal, economic and risk issues.
Industrial Chemicals (GL10.121). SMRI 1998 Spring meeting, Technical Conference, 19–22 April 1998, New Orleans,
Gerritse, J., Van der Grift, B., Langenhoff, A., 2009. Contaminant behaviour of micro- Louisiana, USA.
organisms in groundwater. In: Quevauviller, P., Fouillac, A.M., Grath, J., Ward, R. Verweij, J.M., Simmelink, H.J., 2002. Geodynamic and hydrodynamic evolution of the
(Eds.), Groundwater Monitoring. John Wiley & Sons, Ltd. ISBN: 978-0470-77809-8. Broad Fourteens Basin (The Netherlands) in relation to its petroleum systems. Mar.
Giles, H.N., 1994. The impacts of gas intrusion and geothermal heating on crude oil Pet. Geol. 19, 339–359.
stockpiled in de U.S. Strategic Petroleum Reserve. Presented at the SMRI-Fall meeting White, M.D., Oostrom, M., Lenhard, R.J., 1995. Modeling fluid flow and transport in
1994, Hannover, Germany. variably saturated porous media with the STOMP simulator. 1. Non-volatile three-
Hendriks, D.M.D., Kuijper, M.J.M., Van Ek, R., 2013. Groundwater impact on environmental phase model description. Adv. Water Resour. 18, 53–77.
flow needs of streams in sandy catchments in The Netherlands. J. Hydrol. Sci. Wildenborg, A.F.B., Leijnse, A.L., Kreft, E., Nepveu, M.N., Obdam, A.N.M., Orlic, B., Wipfler,
(accepted). E.L., van der Grift, B., van Kesteren, W., Gaus, I., Czernichowski-Lauriol, I., Torfs, P.,
Jaynes, E.T., 2003. The Theory of Probability, the Logic of Science. Chapter 2 Cambridge Wójcik, R., 2005. Risk assessment methodology for CO2 storage — the scenario
University Press. approach. In: Benson, S.M. (Ed.), The CO2 Capture and Storage Project for Carbon
Kaluarachchi, J.J., Parker, J.C., 1992. Multiphase flow with a simplified model for oil Dioxide Storage in Deep Geological Formations for Climate Change Mitigation.
entrapment. Transp. Porous Media 7, 1–14. Elsevier (Ch. 30).
Knol, A.B., Slottje, P., Van der Sluijs, J.P., Lebret, E., 2010. The use of expert elicitation in Zhang, M., Bachu, S., 2011. Review of integrity of existing wells in relation to CO2
environmental health impact assessment: a seven step procedure. Environ. Heal. 9, 19. geological storage: what do we know? Int. J. Greenh. Gas Control 5, 826–840.

Potrebbero piacerti anche