Вы находитесь на странице: 1из 8

1.

Introduction
Volcanic terranes are the epitome of a dynamic earth system and their
associated hydrothermal systems testify to the importance of water as a vital agent of
heat and chemical transport. These systems exercise control over many volcanic
phenomena, and so understanding their internal workings is important to hazard
forecasting [Houghton et al, 1987; Lopez et al, 1993; Bedrosian et al, 2007]. Fluid
movements can also interfere with the monitoring data we rely on to understand these
terranes, as hydrothermal “scrubbing” and sealing can obscure our view of magmatic
developments [Edmonds et al, 2001 & 2003]. Hence, a knowledge of fluid movement
within volcanic terranes could be key to monitoring system improvements. This
review aims to outline the internal processes of hydrothermal systems, their role as
agents of mass movement beneath volcanoes and their modulation of volcanic
hazards. As we will see, they are responsible for many of the phenomena that make
volcanoes intrinsically unpredictable.
Past research into the dynamics of terrestrial hydrothermal systems has centered
mainly upon ore forming processes. These studies were limited by computational
power and the availability of sufficiently accurate geochemical data. Both of these
factors have improved in recent years, and numerical codes can now couple complex
geochemical calculations with fluid flow simulations on standard personal computers.
These Reactive transport codes are thus able to to tackle complex hydrothermal
studies [Xu et al, 2006; Steefal et al, 2005]. It is this approach that will be employed
by the associated research project upon a case study of the Soufriere Hills Volcano,
Montserrat (hereafter referred to as SHV). The system is one of the most active in the
West Indies [Boudon et al, 1998] and offers an opportunity to model subsurface
processes in a juvenile terrane. This is a well monitored site with a detailed and recent
publication base that will allow collected literature to contribute data towards
simulations. Unfortunately, only two publications detail the SHV hydrothermal
system itself [Chiodini et al, 1996; Boudon et al, 1998] and there is no pre-existing
reference framework for subsurface dynamics. Numerical methods could thus present
an exciting, holistic approach that has already provided unique insights elsewhere
[Todesco et al 2003 & 1995].
2. Hydrothermal systems
Hydrothermal systems develop where a high heat flux and a sufficiently
permeable reservoir (>10-16m2) [Norton et al, 1984; Henley et al, 1983] allow
convective circulation to occur [Norton et al, 1984]. Volcanic terranes supply both of
these requirements by magmatic intrusion. Magmatic presence inevitably dominates
heat flux, but minor contributions are made within the reservoir itself by
devitrification, mineral alteration and radioactive decay [Henley et al, 1983]. The
imposed strain of intrusion can also lead to fracture development and a heterogeneous
reservoir permeability [Norton et al, 1984]. This prevents deep fluids from reaching
equilibrium with shallow-level aquifers on upflow, allowing fumarole (steam vents)
and hot pools formation [Ingebritsen et al, 2006]. Reservoir fluid is principally
meteoric water, circulated to depth by convection. However, near coastlines, hybrid or
seawater dominated systems can occur [Henley et al, 1983; Lachassagne et al, 2009].
Such fluid chemistry may be prevalent at Montserrat, as the volcano occurs only 4 km
from the sea.
As these fluids percolate downward, they encounter pressures that prevent
boiling at their given temperature, and the fluid is kept liquid. At depths of around
500m temperatures may be such that conditions exceed the critical point of water (347
°C and 221 bar), and the fluid is said to be supercritical [Ingebritsen et al, 2006; Robb
et al, 2005](see figure 1a &1b). Liquid/gas phase definitions break down as the
vapour density maximum coincides with the liquid density minimum and the two
phases become physically indistinguishable (As can be seen above the critical point in
figure 1b). In a supercritical state, liquid thermal expansivity is at its maximum,
increasing intermolecular space and reducing stable intermolecular interactions. The
expansion therefore leads to a minimum in kinematic viscosity. This then provides a
high enthalpy fluid with little internal friction; an extremely efficient heat transport
media. The presence of supercritical fluids then leads to high Rayleigh number
convection, or “superconvection”. Because of this mechanism, rates of heat transport
in circulating reservoirs may be 70 times greater than conduction alone [Dunn and
Hardee, 1981]. The movement of these fluids will then have profound implications
for the subsurface circulation and thermal structure [Kawada et at, 2004]. This free
movement of these fluids could permit magmatic perturbations at depth to quickly
move through the hydrothermal network, potentially creating a window upon the
coupled magmatic system.
Upon deep fluid upflow, decompression boiling creates regions where liquid
and vapour fluids coexist in the reservoir lithology and a density contrast occurs
between the two phases. If permeability is available in both horizontal and vertical
directions [Ingebritsen et al, 2006], this density contrast can lead to a modification of
respective upflow vectors. As fluids encounters the regional background
topographically-driven flow, each vector deflects differently depending upon the
respective phase densities and phase separation occurs (see figure 2). On Monserrat,
this phenomenon could contribute to the geographical isolation of a hot spring near
the coast while fumarolic fields occur upon the volcano flanks. Boiling is also a major
control of mineralisation within terrestrial volcanic terranes [Henley et al, 1983] and
this process, in turn, is key to hydrothermal control of magmatic systems [Boudon et
al, 1998, Edmonds et al, 2003, Chiodini et al, 1996].
3. Hydrothermal control of hazards in volcanic terranes
Hydrothermal circulation in volcanic terranes, can only lead to increased
complexity in an already unpredictable system. Many hazards within volcanic
systems are affected by fluid alteration or mineralisation aas these processes directly
modify strength and permeability parameters [Vries et al, 2009; Reid, 2009]. The
occurrence of alteration and mineralisation is common within volcanic terranes,
though it dominantly occurs within the subsurface. Volumes of affected rock may
therefore be difficult to estimate; a concern when assessing hazard risk in volcanic
regions.
Circulating hydrothermal fluids have become increasing linked to volcanic
slope stability in recent years as they chemically alter edifice lithology [Lopez et al,
1993; Kilburn et al, 1998; Reid et al 2004 & 2001; Vries et al, 2009; Boudon et al,
1996]. The susceptibility of a rock to alteration depends primarily upon bulk
mineralogy [Henley et al, 1983], though stress corrosion is an exacerbating factor
[Kilburn et al, 1998]. Volcanic glass is most reactive as it is already metastable;
altering to opal before breaking down further to layered clays [Henley et al, 1983;
Dobson et al, 2004]. These important end-products of alteration permit increased pore
pressures and lower the effective strength of a parent rock. This critically reduces
edifice shear strength [Vries et al, 2009; del Potro et al, 2009], and leads to
catastrophic gravitational slope collapses. If rich in clays, collapse material may
create cohesive debris flows [Vallance et al, 2009]. Such hazards are especially
dangerous and do not require a magmatic trigger [Vries et al, 2009]. At SHV, The
Boxing Day Collapse, 26 December 1997 occured due alteration around a steam-vent
field [Calder et al, MVO Special Report 6]. It is therefore clear that identifying
regions where fluids are at their most reactive, should be an important consideration
when assessing volcanic risk [Bedrosian et al, 2007].
Mineralisation is dominantly controlled by boiling processes. Consequently,
regions where it occurs can become mineralised to an extent where they effectively
seal against fluid flow [Henley et al, 1983; Lachassagne et al, 2009]. Boiling is
strongly pressure dependent, and only occurs when lithostatic pressure is low enough.
Because of this, most mineralisation occurs at shallow levels [Robb et al, 2005]. This
coincides with the exsolution depths of many magmatic volatiles [Boudon et al, 1983]
which rely upon permeable pathways to move through the country rock. Thus,
mineralisation has profound implications for magmatic ascent speeds and volcanic
explosivity [Edmonds et al, 2003].
The transport of mineral saturated fluids can also define the spatial limits of a
hydrothermal system by “sealing”. Upon reaching the cool fringes of the reservoir,
the solubility of most minerals decreases and are precipitated within voids and
fractures, reducing horizontal permeability [Henley et al, 1983]. Mineralisation is the
most effective sealing process in diffuse degassing pathways, though fracture
pathways may be sealed an order of magnitude faster by thermo-expansion [Lowell et
al, 1993]. Fractures allow more efficient degassing but can close by expansion due to
increasing heat flux from magmatic recharge, or supercritical fluid infiltration. In two-
phase regions of flow, severe permeability drops may also occur as fluids compete for
void spaces in the rock [Ingebritsen et al, 2006]. At certain ratios of gas and liquid,
phase “competition” can occur to such an extent that degassing pathways are
effectively blocked [Ingebritsen et al, 2006].
At SHV, mineral sealing is inferred in the shallow aquifer, where a silicified
zone excludes liquid circulation outside of the volcanic crater [Boudon et al, 1998].
Since the beginning of activity in 1995, it is though that the conduit region has dried
out and no groundwater reaches this region [Druitt et al, 2002]. Instead,
mineralisation occurs by vapor transport [Boudon et al, 1998], a poorly understood
realm of geochemistry that is invoked in ore forming processes. At Hakone, Japan,
mineralisation prevents horizontal intrusion of seawater into the system [Henley et al,
1983]. As SHV is only 4km from the coast, it is possible the same could occur here.
Vertically, such sealing is rare as flow paths to the surface are maintained actively by
hydrofracturing [Henley et al, 1983]. In dramatic cases, this process leads to
hydrothermal explosions or “steam-blasts” [Barberi et al, 1992].
These “steam-blasts” (terminology Jaggar 1949) are evidence of disequilibrium
between magmatic and hydrothermal systems [Savin 2005, Boudon et al, 1998]. They
pose a major threat to surrounding populations [Houghton et al, 1987] and are
associated with ballistics, deadly gases, lahars, base-surges and debris flows [Barberi
et al, 1992 and refs therin]. They have two main triggers; pressure-release or heat-
increase, both of which are inevitable in a juvenile volcanic terrane. Overpressure
release by seismic or hydraulic fracturing [Savin et al, 2005] or sudden unloading,
can result in a sufficient lithostatic pressure drop to allow catastrophic vapor
expansion. This vapor expansion is also seen by increasing the shallow-level heat flux
[Savin 2005], which can occur by magma intrusion or by superheated fluid
introduction. The latter case is especially relevant at SHV as magmatically-derived
fluids have access to shallow level reservoirs [Hammouya et al, 1998; Chiodini et al,
1996]
4. Volcanic monitoring in hydrothermally active terranes
As volcanic systems become more complex with circulating fluids presence, so
too does the task of hazard monitoring. Seismic, deformation and gas emission
surveys are key methods in volcanic monitoring but hydrothermal “noise” can be a
problem. Understanding this “noise” may be important for the advancement of hazard
monitoring and our knowledge of subsurface hydrothermal processes.
“Steam-blasts” have been reported to exhibit seismic precursors that can be
confused with magmatic siesmicity [Savin et al, 2005; Barberi et al, 1992; Druitt et al,
2002]. However this is not always the case, and many steam blasts occur unheralded,
especially if seismic arrays center on deeper perturbations [Barberi et al, 1992].
Seismic monitoring is also subject to noise from shallow rock failure under gravity
[Moran et al, 2000]. This is a less hazardous manifestation of the same alteration
process that leads to massive edifice collapses and warns of gradual edifice
weakening. Ground deformation studies may also suffer from hydrothermal
interference as fumarolic venting can cause hydrothermal reservoir deflation, this has
been reported at at Kuju Volcano, Japan [Ehara et al, 2005]. Here, the resulting
contraction across the edifice could have been misinterpreted as subsidence due to
magmatic quiescence. In a gas monitoring studies of SHV, Edmonds et al, 2003
inferred hydrothermal permeability modification within the edifice and dome as the
main control of short term SO2 fluctuations (see figure 3). Whilst an insightful and
pioneering approach to the use of gas data, we do lose the direct monitoring ability of
this method. If numerical models could constrain edifice and dome permeabilities,
then SO2 methods could again be used on shorter time scale studies.
5. Numerical modeling
Numerical models enable an exciting new approach to earth systems science.
While research of fundamental processes will always remain crucial to advancement
of our understanding, broader studies allow a contextual evaluation of such processes.
As we have seen, the dynamics of volcanic regions are subject to many competing
mechanical and chemical processes and numerical investigation can provide an
important interpretive tool [Steefal et al, 2005]. The ability to vary system parameters
allows quantitative study of situational evolution and lets us “pick-apart” these
complex systems process by process. We can then evaluate the importance of each
system property. This ability can facilitate a more in-depth understanding of how
dynamic processes interact with situational evolution. Such knowledge is important if
we are to hope for more accurate hazard forecasting methods.
Micol Todesco, who has studied a number of prominent Italian with these
methods, most notably Vulcano and Phlegrean Fields caldera. Using Lawrence
Berkeley’s TOUGH2 code, the thermal and circulation structure was inferred beneath
Vulcano in the Aeolian Islands [Todesco 1995]. The model used pure H2O heat
transport without any invoked chemical reactions. It revealed not only the generalised
flow in the subsurface but allowed the permeability distribution within the subsurface
to be inferred. When studying Phlegrean Fields caldera, near Naples, Todesco
coupled TOUGH2 with a mechanical stress code, FLAC, to create TOUGH-FLAC
[Todesco 2003a]. The model revealed that regional heave and deformation was due to
fluid boiling within the aquifer and not due to magmatic recharge beneath the caldera.
Such results show the power of applied numerical modeling.
TOUGHREACT is the newest code from Lawrence Berkeley national
laboratory and will be employed by this research project to model the hydrothermal
system of SHV. It is a reactive transport code that runs geochemical and fluid flow
codes in tandem [Xu et al 2000 & 2006]. It has been used for CO2 sequestration
studies, geothermal reservoir modeling [Xu et al 2006] and has proven successful in
work on Yellowstone hydrothermal system. Here geochemical simulations discovered
that the system fluid chemistry was dominanted by equilibria with one volumetrically
small rock unit [Dobson 2004]. Though TOUGHREACT has a record of success, it
does have limitations. The Lawrence Livermore geochemical dataset used by
TOUGHREACT is only accurate up to temperatures of 300 °C and hence cannot
model supercritical chemical behaviour. Any model produced with this code can
however, model deep reservoir heat transport. With these results, we can then define
heat flow near the surface for shallow-level geochemistry. The research scope is then
unaltered, as most important hazards develop in shallow regions where magma is in
greater disequilibrium [Boudon et al, 1998].
6. Case Study: The Soufriere hills Hydrothermal System
SHV possesses one of the most active hydrothermal systems on the Lesser
Antilles arc [Boudon et al, 1998]. It is well expressed at the surface by four large
fumarolic fields [Chiodini et al, 1996] and hot springs near the west coast (see figure
4 and 5b). The fumarolic fields, in order of size are are Galway’s Soufriere, the Gages
Soufrieres (Upper and Lower) and The Tar River Soufriere [Chiodini et al, 1996].
Galway’s Soufriere contains hundreds of steam vents and several boiling pools and
has been a key interest to previous works [Boudon, Chiodini]. All of the soufrieres are
associated with large deposits of amorphous silica, clays and alunite [Chiodini et al,
1996]; displaying strong alteration and mineralisation. Such processes have already
been linked to many hazards at SHV, most prominently the 2006 boxing day collapse
[Calder et al, MVO Special Report 6]. Hence, continued oversight of these processes
is a extremely dangerous.
The subsurface system itself was linked to magmatic activity in the early 1990s,
before recent eruption activity began [Chiodini et al, 1996]. Thermal springs on the
west coast, near Plymouth, and fumarolic fields upon the volcano edifice were shown
to exhibit magmatic 3He/4He ratios [Hammouya, 1998; Chiodini et al, 1996].
Galway’s Soufriere displayed an especially strong 3He/4He signal, suggesting it is
located on a deep penetrating, high permeability zone. This feature may significantly
control hydrothermal structure, by allowing thermal and chemical movement between
different reservoirs in the system [Kawada et al, 2004]. The deep reservoir
temperatures were also inferred to be ~250 °C; an important constraint for the deep
source of the fumarolic fields [Chiodini et al, 1998]. These conclusions from Chiodini
et al, 1996 and Boudon et al, 1998, have been used to develop the subsurface
conceptual model in figure 5.
7. The future
This review has highlighted a realm that is relatively unexplored in the science
of volcanology. While hazard forecasting has mainly progresses by improving our
understanding of magmatic processes [Sparks 2003], we could greatly benefit from a
wider understanding of terrane dynamics. This could be seen as a deductive, or
holistic approach to volcanology and could accelerate our advancement by removing
the “noise” experience by magmatic monitoring systems. Hydrothermal systems are a
major source of monitoring “noise” in volcanic terranes and exhibit seismicity and
ground deformation which can be mistaken for magmatic activity. Removing such
confounding factors from our monitoring data will allow a clearer picture of
magmatic evolution beneath volcanic terranes.
Despite their importance, hydrothermal systems are quite poorly understood.
Heat transfer dynamics are poorly constrained [Lachassagne et al, 2009; Ingebritsen
et al, 2006], reducing accuracy in thermal structure inferences. Publications rarely
detail aquifer structures or flow direction and any attempts generally lack quantitative
constraints. The nature of numerical modeling addresses these deficiencies and can
help put new constraints upon system dynamics. Sensitivity analyses can identify key
features of a system supported by recorded field data, maintaining important
contextual constraints. Unfortunately, heterogeneous permeability is still difficult to
identify by field data and when unidentified, can cause considerable inaccuracy in
model systems. The remedy to this is a selection of new methods to explore the
subsurface permeability structure. Voids and fractures can now be imaged using
seismic tomography, high resolution reflection surveys [Gritto et al, 2002 & 2003] or
inferred from seismic anisotropy studies. The latter is already yielding results in oil
reservoir flow models [Maultsch et al, 2003; Rogers 2009] and raises hopes that
numerical modeling could soon become an integral part of volcanological study.
References
Barberi F, Bertagnini A, Landi P, Principe C. (1992) A Review of Phreatic
Eruptions and Their Precursors. Journal Of Volcanology And Geothermal Research
vol. 52 pp. 231-246
Barnes, L. (2005) Introduction to Ore Forming Processes. USA, UK, Australia.
Blackwells Publishing.
Bedrosian PA, Unsworth MJ, Johnston MJS. (2007) Hydrothermal circulation
at Mount St. Helens determined by self-potential measurements. Journal Of
Volcanology and Geothermal Research, vol. 160 (1-2) pp. 137-146
Boudon G, Villemant B, Komorowski JC, Ildefonse P, Semet MP. (1998)
The hydrothermal system at SHV, Montserrat (West Indies): Characterization and
role in the on-going eruption. Geophysical Research Letters vol. 25 pp. 3693-
3696
Calder et al. 1996. MVO Special Report 6. The Boxing Day Collapse
http://www.geo.mtu.edu/volcanoes/west.indies/soufriere/govt/specrep/specrep06.
html
Chiodini G, Cioni R, Frullani A, Guidi M. (1996) Fluid geochemistry of
Montserrat Island, West Indies. Bulletin of Volcanology. Vol. 58 pp.380–392
de Vries BV, Kerle N, Petley D. (2000) Sector collapse forming at Casita
volcano, Nicaragua. Geology vol. 28 pp. 167-170
del Potro and Huerlimann M. (2009) The decrease in the shear strength of
volcanic materials with argillic hydrothermal alteration, insights from the summit
region of Teide stratovolcano, Tenerife. Engineering Geology vol. 104 pp. 135-143
Dobson PF, Salah S, Spycher N. (2004) Simulation of water–rock interaction
in the Yellowstone geothermal system using TOUGHREACT. Geothermics
Dunn, J. C. and Hardee H. C. (1981) Superconvecting geothermal zones.
Journal of Volcano & Geothermal Research. Vol. 11, pp. 189 –201
Druitt TH, Young SR, Baptie B, Bonadonna C. (2002) Episodes of cyclic
Vulcanian explosive activity with fountain collapse at Soufriere Hills Volcano,
Montserrat. Geological Society London Memoirs
Edmonds M, Pyle DM, Oppenheimer C. (2001) A model for degassing at the
SHV, Montserrat, West Indies, based on geochemical data. Earth And Planetary
Science Letters. Vol. 186, pp.159-173.
Edmonds M, Oppenheimer C, Pyle DM, Herd RA. (2003) SO2 emissions from
Soufrière Hills Volcano and their relationship to conduit permeability,
hydrothermal interaction and degassing regime. Journal Of Volcanology And
Geothermal Research Vol 124, pp. 23-43.
Ehara S, Fujimitsu Y, Nishijima J, Fukuoka K. (2005). Change in the Thermal
State in a Volcanic Geothermal Reservoir beneath an Active Fumarolic Field after
the 1995 Phreatic Eruption of Kuju Volcano, Japan. Proceedings of the World
Geothermal Congress, Antalya, Turkey, 24-29 April 2005
Elsworth D, Voight B, Thompson G, Young SR (2004). Thermal-hydrologic
mechanism for rainfall-triggered collapse of lava domes. Geology Vol 32, pp. 969-
972.
Fujimitsu. Y; Ehara, S; Oki, R; Kanou, R. (2008) Numerical model of the
hydrothermal system beneath Unzen Volcano, Japan. Journal Of Volcanology And
Geothermal Research vol. 175, pp. 35-44
Gritto, R; Daley T. M; Majer, E. L. (2002) Integrated seismic studies at the
Rye Patch Geothermal Reservoir, Nevada. Conference: Geothermal Resources
Council Annual Meeting 2002.
Gritto, R. (2003) Subsurface void detection using seismic tomographic imaging.
Conference: Geothermal Resources Council Annual Meeting 2003.
Hammouya G, Allard P, Jean-Baptiste P, Parello F, Semet MP, Young SR.
(1998) Pre- and syn-eruptive geochemistry of volcanic gases from SHV of
Montserrat, West Indies. Geophysical Research Letters vol. 25 pp. 3685-3688
Henley RW, Ellis AJ. (1983) Geothermal systems ancient and modern: a
geochemical review. Earth Science Reviews Vol. 19, pp. 1--50.
Houghton BF, Latter JH, Hackett WR. (1987). Volcanic hazard assessment for
Ruapehu composite volcano, Taupo volcanic zone, New Zealand. Bulletin of
Volcanology Vol.49 pp. 737-751.
Ingebritsen, S. Sanford, W. Neuzil, C (2006) Groundwater in Geologic
Processes (Second Edition), USA: Cambridge University Press.
Jaggar T.A., 1949. Steam blast volcanic eruptions. 4th Special Report of the
Hawaiian volcano observatory.
U.S. Geol. Surv. and the Hawaiian Volcano Research Association.
Kawada Y, Yoshida S, Watanabe S. (2004) Numerical simulations of mid-
ocean ridge hydrothermal circulation including the phase separation of seawater.
Earth Planets Space, 56, 193–215
Kilburn et al, & Voight (1998). Slow rock fracture as eruption precursor at
Soufriere Hills Volcano, Monsterrat. Geophysical Research Letters vol. 25, NO. 19,
Pages 3665-3668,
Lachassagne P, Marechal J, Sanjuan B, (2009). Hydrogeological model of a
high-energy geothermal field (Bouillante area, Guadeloupe, French West Indies).
Hydrogeology Journal vol. 17 pp. 1589-1606
Lopez DL, Williams SN (1993). Catastrophic volcanic collapse: relation to
hydrothermal processes. Science Vol 260, pp. 1794-1796
Lowell RP, Cappellen PV, Germanovich LN. (1993) Silica Precipitation in
Fractures and the Evolution of Permeability in Hydrothermal Upflow Zones.
Science vol. 260 pp. 192-194
Maultzsch S, Chapman M, Liu ER, Li X.Y. (2003) Modelling frequency-
dependent seismic anisotropy in fluid-saturated rock with aligned fractures:
implication of fracture size estimation from anisotropic measurements.
Geophysical Prospecting vol. 51 pp. 381-392
Moran SC, Zimbelman DR, Malone SD. (2000). A model for the magmatic–
hydrothermal system at Mount Rainier, Washington, from seismic and
geochemical observations. Bulletin of Volcanology Vol. 61. pp. 425-436.
Norton et al,. (1984) Theory of Hydrothermal Systems. Annual Review of Earth
and Planetary Sciences vol. 12 pp. 155-177
Reid ME, Sisson TW, Brien DL. (2001) Volcano collapse promoted by
hydrothermal alteration and edifice shape, Mount Rainier, Washington. Geology
vol. 29 pp. 779-782
Reid. (2004) Massive collapse of volcano edifices triggered by hydrothermal
pressurization. Geology vol. 32 pp. 373-376
Robb, H (1997) Geochemistry of Hydrothermal Ore Deposits (Third Edition),
USA: John Wiley & Sons, Inc.
Rogers, SLE, 2009. Constraining models of fractured reservoirs using seismic
anisotropy maps, for improved
reservoir performance and prediction. British Geological Survey: Edinburgh
Anisotropy Project
Sparks, RSJ. (2003) Forecasting volcanic eruptions. Earth & Planetary Science
Letters vol. 210 pp. 1-15
Steefel CI, DePaolo DJ, Lichtner PC. (2005) Reactive transport modeling: An
essential tool and a new research approach for the Earth Sciences. Earth And
Planetary Science Letters, Vol 240, pp. 539-558.
Todesco M, Rutqvist J, Pruess K, Oldenburg C. (2003) Multi-phase fluid
circulation and ground deformation: a new perspective on bradyseismic activity at
Phlegrean Fields. Proc. 28th Workshop on Geothermal Reservoir Eng.
Todesco, M. (1995) Modeling of the geothermal activity at Vulcano (Aeolian
Islands, Italy). Proceedings of the World Geothermal Congress 1995.
Vallance JW, Scott KM. (1997). The Osceola Mudflow from Mount Rainier:
Sedimentology and hazard implications of a huge clay-rich debris flow. Geological
Society Of America Bulletin Vol. 109, pp. 143-163
Xu, T; Pruess K. (2000) Multiphase fluid flow and subsequent geochemical
transport in variably saturated fractured rocks: Lawrence Berkeley National
Laboratory supporting documents of TOUGHREACT.
Xu T, Sonnenthal E, Spycher N, Pruess K. (2006) TOUGHREACT—A
simulation program for non-isothermal multiphase reactive geochemical transport
in variably saturated geologic media: Applications to geothermal injectivity and
CO2 geological sequestration. Computers and Geosciences Vol. 32, pp. 145-165.

Вам также может понравиться