Earth:Fault zone hydrogeology

From HandWiki

Fault zone hydrogeology is the study of how brittlely deformed rocks alter fluid flows in different lithological settings, such as clastic, igneous and carbonate rocks.[1] Fluid movements, that can be quantified as permeability, can be facilitated or impeded due to the existence of a fault zone.[1] This is because different mechanisms that deform rocks can alter porosity and permeability within a fault zone.[1][2] Fluids involved in a fault system generally are groundwater (fresh and marine waters) and hydrocarbons (Oil and Gas).[3]

Take notice that permeability (k) and hydraulic conductivity (K) are used interchangeably in this article for simplified understanding

GIF 1. This GIF shows how fault zones affect fluid migrations in the cross section view. A) The fault zone acts as a barrier that blocks fluid flows across it. B) The fault zone acts as a conduit that allows or facilitates fluid flows across it.

Architecture

Figure 1. The figure shows the architecture of a fault zone, in which a fault core is enveloped by a damage zone.

A fault zone can be generally subdivided into two major sections, including a Fault Core (FC) and a Damage Zone (DZ) [4][5] (Figure 1).

The fault core is surrounded by the damage zone. It has a measurable thickness which increases with fault throw and displacement, i.e. increasing deformations.[1]

The damage zone envelopes the fault core irregularly in a 3D manner which can be meters to few hundred meters wide (perpendicular to the fault zone).[6] Within a large fault system, multiple fault cores and damage zones can be found.[1] Younger fault cores and damage zones can overlap the older ones.

Different processes can alter the permeability of the fault zone in the fault core and the damage zone will be discussed respectively in the next section. In general, the permeability of a damage zone is several orders of magnitude higher than that of a fault core as damage zones typically act as conduits (will be discussed in section 3).[7] Within a damage zone, permeability decreases further away from a fault core.[7]

Permeability classification

There are many classifications to group fault zones based on their permeability patterns. Some terms are interchangeable; while some have different subgroups. Most of expressions are listed in the following table for comparison. Dickerson's categorisation is commonly used and easier to understand in a broad range of studies.[4]

The classification of a fault zone can change spatially and temporally. The fault core and damage zone can behave differently to accommodate the deformations.[1] Moreover, the fault zone can be dynamic through time. Thus, the permeability patterns can change for short-term and long-term effects.[1]

Table 1 Common Permeability Classification of Fault Zone
Authors / Definitions Kfz < Khr Kfz < Khr and Kfz > Khr Dynamic System Kfz > Khr
Dickerson's (2000) [8] Barriers Barrier-Conduit / Conduit
Aydin's (2000) [9] Transmitting Vertically Transmitting and Lateral Sealing Transmitting or Lateral Intermittently Sealing
Caine & Foster's (1999) [10] Quantitative mapping and discrete fracture models
Others [4] Detailed description (Weak/ Strong); Barrier/ Conduit Permeability Ratio

*K = Permeability/ Hydraulic Conductivity

*fz = fault zone

*hr = host rock = Undeformed rock surrounds the fault zones

Mechanisms (permeability)

Figure 2. The figure shows a cross-section that consists of a fault zone cutting through the sandstone and shale layers. A zoom-in box illustrates the sand-sized grains within the sandstones.

Fault zone results from brittle deformation.[3] Numerous mechanisms can vary the permeability of a fault zone. Some processes affect the permeability temporarily. These processes enhance the permeability for a certain period, and then reduce it later on: in this case, like seismic events, the permeability is not constant through time.[11] Physical and chemical reactions are the major types of mechanisms.[1] Different mechanisms can occur different in fault core and damage zone since the intensities of deformation they experience are different (Table 3).

Table 2 Occurrences in Fault Core and Damage Zone [1][12]
Mechanisms Deformation Bands Brecciation Fracturing Sediment Mixing Clay Smears Cataclasis Compaction Dissolution Precipitation Seismic Event
Fault Core + + + + + + + + +
Damage Zone + + + +

*+ = more likely to occur at

Enhancing fault zone permeability

Deformation bands

Figure 3. The figure shows a dilation band. Its formation does not involve grain movements and it can facilitate fluid movements.

The formation of a dilation band, in unconsolidated material, is the early result of applying extensional forces.[1] Disaggregation of mineral fabric occurs along with the band, yet no offset is accommodated by the movement of grains [1] (Figure 3).

Figure 4. This shows how a shear band facilitates fluid movements by rotation and sliding.

Further deformation causes offsets of mineral grains by rotation and sliding.[3] This is called a shear band. The pore network is rearranged by granular movements (also called particulate flow), hence moderately enhance permeability. However, continuing deformation leads to the cataclasis of mineral grains which will further reduce permeability later on (section 3.2.3) [1] (Figure 4).

Brecciation

Figure 5. This shows the brecciation of rocks, thus, enhancing permeability by opening new paths.

Brecciation refers to the formation of angular, coarse-grained fragments embedded in a fine-grained matrix.[13] As breccia (the rock experienced brecciation) is often non-cohesive, thus, permeability can be increased up to four or five orders of magnitude.[1] However, the void space enlarged by brecciation will lead to further displacement along the fault zone by cementation, resulting in a strong permeability reduction [1] (Figure 5).

Fracturing

Figure 6. The figure shows the formation of a fracture, providing a relatively large aperture for fluid flow.

Fractures propagate along a fault zone in direction responding to the stress applied.[5] The enhancement of permeability is controlled by the density, orientation, length distribution, aperture, and connectivity of fractures.[4] Even fracture with aperture of a 100-250 μm can still greatly influence fluid movement (Figure 6).[1]

Reducing fault zone permeability

Sediment mixing

Sediments, typically from distinct formations, with different grain sizes, are mixed physically by deformation, resulting in a more poorly-sorted mixture. The pore space is filled by smaller grains, increasing tortuosity (mineral scale in this case) of fluid flow across the fault system.[1]

Clay smears

Figure 7. The figure shows clay smears formed by deformation within a fault zone, providing a sealing effect for fluids.

Clay minerals are phyllo-silicate, in other words, with sheet-like structure.[14] They are effective agents that block fluid flows across a fault zone.[14] Clay smears, deformed layers of clay, that are developed along fault zone can act as a seal of hydrocarbon reservoir i.e. extremely low permeability that nearly prohibits all fluid flows (Figure 7).[1]

Cataclasis

Cataclasis refers to pervasive brittle fracturing and comminution of grains.[15] This mechanism becomes dominant at depth, greater than 1 km, and with larger grains.[1] With increasing intensity of cataclasis, fault gouge, often with the presence of clay, is formed.[1] The largest reduction occurs on the flows that are perpendicular to the band.[1]

Enhancing and reducing fault zone permeability successively

Compaction and cementation

Compactions and cementation generally lead to permeability reduction by losing porosity.[1] When a large region, which consist a fault zone, experience compaction and cementation, porosity loss in host rock (undeformed rock surrounding the fault zone) can be greater than that of fault zone rock. Hence, fluids are forced to flow through a fault zone.[1]

Dissolution and precipitation

Solute carried by fluids can either enhance or reduce permeability by dissolution or precipitation (cementation).[1][16] Which process takes place depends on geochemical conditions like rock composition, solute concentration, temperature, and so on.[1] The changes in porosity dominantly control whether the fluid-rock interaction continues or slows down as a strong feedback reaction.

For example, minerals like carbonates, quartz, and feldspars are dissolved by the fluid-rock interactions due to enhanced permeability.[1] Further introduction of fluids can either continuously dissolve or otherwise re-precipitate minerals in the fault core, and thus alters the permeability.[1] Therefore, whether the feedback is positive or negative heavily depends on the geochemical conditions.

Seismic event

Earthquakes can either increase or decrease permeability along fault zones, depending on the hydrogeological settings. Recorded hot spring discharges show seismic waves dominantly enhance permeability,[11][17][18] but reductions in discharge may also result occasionally.[19] The timescale of the changes can be up to thousands of years.[16] Hydraulic fracturing (fracking) requires increasing the interconnectedness of the pore space (in other words, permeability) of shale to allow the gas to flow through the rock, and very small deliberately induced seismic activity of magnitudes smaller than 1 are applied to enhance rock permeability.[20]

Taking the Chile earthquake in 2017 as an example, the discharge of streamflow temporally increased six times indicates a sixfold enhancement in permeability along the fault zone.[11] Yet, seismic-induced effects are temporary that normally last for months, for Chile's case, it lasted for one and a half months which gradually decreased back to original discharge.[11]

Table 3 Fault Zone Permeability Changes [1]
Mechanisms Dilation Shear Bands Brecciation Fracturing Sediment

Mixing

Clay Smears Cataclasis

(with low presence of clay)

Changes in Order(s) of Magnitude

(compared to no fault zone)

1 < 1 4 ~ 5 2 ~ 3 - 2 Act as a seal [21] - 1 ~ - 3

(- 4 ~ - 5)

Mechanisms (porosity)

Porosity (φ) directly reflects the specific storage of rock. And brittle formation alters the pores by different mechanisms. If the pores are deformed and connected together, the permeability of rock enhances.[2] On the other hand, if the deformed pores disconnect with each other, the permeability of rock in this case reduces.[2]

Pore types

Table 4 Four Major Pore Types (that are related to fault zone)
Pore types Illustrations Mechanisms Descriptions
Intergranular pores Int Pore.png

Intφ refers to intergranular pores.

Image under petrographic microscope is extracted from Farrel & Healy, 2017.[2]

Compaction

Cementation

Intergranular pores are the voids formed during incomplete cementation.[22] The pore spaces are not occupied by grains. The main feature of intergranular pores is that they are irregular in shape.[23]
Grain dissolution pores Dis Pore.png

Disφ refers to grain dissolution pores.

Image under petrographic microscope is extracted from Farrel & Healy, 2017.[2]

Dissolution Grain dissolution pores are formed when minerals are dissolved out. The pores are left as evidence of pre-existing mineral grains.[24] The minerals usually are feldspars, calcites and so on.[23]
Intragranular fracture pores Ft Pore.png

Ftφ refers to Intragranular fracture pores.

Image under petrographic microscope is extracted from Farrel & Healy, 2017.[2]

Fracturing

Brecciation

Cataclasis

????

Intragranular fracture pores are the voids created within a mineral grain.[22] The fracture can cut through a grain into two grains.
Transgranular fracture pores T-Ft Pore.png

T-Ftφ refers to Transgranular fracture pores.

Image under petrographic microscope is extracted from Farrel & Healy, 2017.[2]

Fracturing

Brecciation

Cataclasis

Transgranular fracture pores are the cracks that cut through more than one mineral grain.[22]

Enhancing porosity

Dissolution

The mineral grains can be dissolved when there is fluid flow. The spaces originally occupied by the minerals will be spare as voids, increasing the porosity of rock.[2] The minerals that are usually dissolved are feldspar, calcite and quartz.[1] Grain dissolution pores results from this process can enhance porosity.

Reducing porosity

Cataclasis, fracturing and brecciation

The mineral grains are broken up into smaller pieces by faulting event. Those smaller fragments will re-organise and further be compacted to form smaller pore spaces.[2] These processes create intragranular fracture pores and transgranular fracture pores.

Important to be aware is that reducing porosity does not equal to reduction in permeability. Fracturing, bracciation and initial stage of cataclasis can connect pore spaces by cracks and dilation bands, increasing permeability.[2]

Precipitation

The mineral grains can be precipitated when there is fluid flow. The voids in the rocks can be occupied by precipitation of mineral grains. The minerals fill the voids and hence, reducing the porosity.[1] Overgrowth, precipitation around an existing mineral grain, of quartz are common.[2] And overgrown minerals infill pre-existing pores, reducing porosity.[2]

Clay deposition

Clay minerals are phyllo-silicate, in other words, with sheet-like structure.[14] They are effective agents that block fluid flows. Kaolinite which is altered from potassium feldspar with the presence of water is a common mineral that fills pore spaces.[2] Precipitation and infiltration only affect materials on shallow depth, hence, more clay materials infill pore spaces when they are closer to the surface. Yet, development of a fault zone introduces fluid to flow deeper.[2] Thus, this facilitates clay deposition at depth, reducing porosity.

Lithological effects

Lithology has a dominant effect on controlling which mechanisms would take place along a fault zone, hence, changing the porosity and permeability.[1]

Table 5 Prominent Lithological Effects on Fault Zone Permeability [1][11]
Mechanism / Lithology Clastic Rocks Igneous Rocks Carbonate Rocks
Deformation Bands
Brecciations
Fracturing
Sediment Mixing
Shale Formation
Clay Smears
Cataclasis
Compaction & cementation
Dissolution
Precipitation
Seismic Events

*↑ = mechanism that enhance permeability

*↓ = mechanism that reduce permeability

Fault type effects

All faults can be classified into three types. They are normal fault, reverse fault (thrust fault) and strike-slip fault. These different faulting behaviours accommodate the displacement in distinct structural ways.

The differences in faulting motions might favour or disfavour certain permeability altering mechanisms to occur.[1] However, the main controlling factor of the permeability is the rock type.[1] Since the characteristics of rock control how a fault zone can be developed and how fluids can move. For instance, sandstone generally has a higher porosity than that of shale. A deformed sandstone in three different faulting systems should have a higher specific storage, hence permeability, than that of shale. Similar example like the strength (resistance to deform) also significantly depends on rock types instead of fault types. Thus, the geological features of rock involved in a fault zone is a more dominated factor.[1][2]

On the other hand, the type of fault might not be a dominant factor but the intensity of deformation is.[1][6] The higher intensity of stresses applied to the rock, the more intense the rock will be deformed. The rock will experience a greater permeability changing event. Thus, the amount of stress applied matters.

Equally important is that identifying the permeability category of the fault zones (barriers, barrier-conduits and conduits) is the main scope of study.[1] In other words, how the fault zones behave when there are fluids pass through.

Studying approaches and methods

Surface and subsurface test

The studies of fault zones are recognised in the discipline of Structural Geology as it involves how rocks were deformed; while investigations of fluid movements are grouped in the field of Hydrology.[1][4] There are mainly two types of methods used to examine the fault zone by structural geologists and hydrologists (Figure 7).

Table 6 Comparisons between Two Major Methods [1][3][4]
Methods Techniques Advantages Limitations
In Situ/ Subsurface-focused test

(Hydrologists)

  • Can study over a vast region
  • Hydraulic properties at depth can be obtained
  • Expensive and inefficient to study in a small-scaled region
  • Limited lithological and structural information
  • Cannot discover the existence of a fault zone at once
Outcrop Mapping/ Surface-focused test

(Structural Geologist)

  • Details of fault zone architectures
  • Less expensive
  • Uncertainties resulted from predictions for rocks at depth
  • Cannot be conducted over a vast region
Figure 8. The schematic diagram shows the approach differences used between Hydrologist and Structural Geologist i.e. subsurface vs surface methods.

In situ test includes obtaining data from boreholes, drill cores, and tunnel projects.[1] Normally the existence of a fault zone is found as different hydraulic properties are measured across it as fault zones are rarely drilled through (except for tunnel projects) (Figure 8).

The hydraulic properties of rocks are either obtained directly from outcrop samples or shallow probe holes/ trial pits,[1] then the predictions of fault structure are made for the rocks at depth (Figure 8).

Example of a subsurface test

An example of a large-scale aquifer test conducted by Hadley (2020), the author used 5 wells aligned perpendicular to the Sandwich Fault Zone in the US, and the drawdowns as well as the recovery rates of water levels in every well were observed.[3] From the evidence that the recovery rates are slower for the wells closer to the fault zone, it is suggested that the fault zone acts as a barrier for northward groundwater movement, affecting freshwater supply in the north.[3]

Example of a surface test

From an outcrop study of the Zuccale Fault in Italy by Musumeci (2015), the surface outcrop findings and cross-cutting relationship are used to determine the number and mechanism of deformational events happened in the region.[25] Moreover, the presences of breccias and cataclasites, that formed under brittle deformation,[25] suggest that there was an initial stage of permeability increase, promoting an influx of CO
2
-rich hydrous fluids.[26] The fluids triggered low-grade metamorphism and dissolution-and-precipitation (i.e. pressure-solution) in mineral scale that shaped a foliated fault core, hence, enhancing the sealing effect significantly.[26]

Other methods

Geophysics

Underground fluids, particularly groundwater, create anomalies for superconducting gravity data which help study the fault zone at depth.[27] The method combines gravitational data and groundwater conditions to determine not only the permeability of a fault zone but also whether the fault zone is active or not.[27]

Geochemistry

The geochemistry conditions of mineral fluids, water or gases, can be used to determine the existence of a fault zone by comparing the geochemistry of fluids' source, given that the conditions of aquifers are known.[28] The fluids can be categorized by the concentrations of common solutes like total dissolved solids (TDS), Mg-Ca-Na/K phase, SO4-HCO3-Cl phase, and other dissolved trace elements.[28][29]

Existing biases

The selection of an appropriate studying approach(es) is essential as there are biases existed when determining the fault zone permeability structure.[4]

In crystalline rocks, the subsurface-focused investigations favor the discoveries of a conduit fault zone pattern; while the surface methods favor a combined barrier-conduit fault zone structure.[4] The same biases, to a lesser extend, exist in sedimentary rock as well.[4]

The biases can be related to the differences in studying scale. For structural geologists, it is very difficult to conduct outcrop study over a vast region; likewise, for hydrologists, it is expensive and ineffective to shorten borehole intervals for testing.[4]

Economic geology

It is economically worth studying the complex system, especially for arid/ semi-arid regions,[30] where freshwater resources are limited, and potential areas with hydrocarbon storages.[1][3][21] Further research on the fault zone, the result of deformation, gave insights on the interactions between earthquakes and hydrothermal fluids along fault zone.[16][27] Moreover, hydrothermal fluids associated with the fault zone also provide information on how ore deposits accumulated.[16]

Artificial hydrocarbon reservoirs

Figure 9. This shows the injection of atmospheric carbon by well, encountering a micro-fracture and a fault zone at depth.

Carbon sequestration is a modern method dealing with atmospheric carbon. One of the methods is pumping atmospheric carbon to specific depleted oil and gas reservoirs at depth. However, the presence of a fault zone act as either a seal or a conduit,[21] affecting the efficiency of hydrocarbon formation.

Micro-fractures that cut along the sealing unit and the reservoir rock can greatly affect the hydrocarbon migration.[21] The deformation band blocks the lateral (horizontal) flow of CO
2
and the sealing unit keeps the CO
2
from vertical migration [21] (Gif 1). The propagation of a micro-fracture that cuts through a sealing unit, instead of having a deformation band within the sealing unit, facilitates upward CO
2
migration (Gif 2). This allows fluid migrations from one reservoir to another.[21] In this case, the deformation band still does not facilitate lateral (horizontal) fluid flow.[21] This might lead to the loss of injected atmospheric carbon, lowering the efficiency of carbon sequestration.

A fault zone that displaces sealing units and reservoir rocks can act as a conduit for hydrocarbon migration.[6] The fault zone itself has higher storage capacity (specific capacity) than that of the reservoir rocks, therefore, before the migration to other units, the fault zone has to be fully filled [6] (Gif 3). This can slower and concentrate the fluid migration. The fault zone facilitates vertical downwards movement of CO
2
due to its buoyancy and piezometric head differences, i.e. pressure/ hydraulic head is greater at a higher elevation, which helps store CO
2
at depth.[6]

Gif 4. This gif shows how fluid (CO
2
) is facilitated by a fault zone. The fault zone acts as a conduit for fluid and it allows fluid migration to lower layers after it is filled with fluid.[6] (One-dimensional movement is illustrated for simplicity)

thumb|left|310x310px|Gif 2. This gif shows how the fluid (CO
2
) is blocked by the deformation band when fluid travels across but not along with it.[21] (One-dimensional movement is illustrated for simplicity)

Gif 3. This gif shows how the fluid (CO
2
) is facilitated by the micro-fracture, which cuts through a sealing unit, within a fault zone. This allows fluid migration to different layers which initially is prohibited by sealing units (shale).[21] (One-dimensional movement is illustrated for simplicity)
Gif 5. The gif shows how fault, induced by seismic events, propagates into a confined aquifer.[31] And fluid of the confined aquifer fills the fault zone and precipitates minerals.[31] Mineralization blocks further fluid movement. Repeated seismic events can deposit economically vulnerable vein structured ore deposit.

Seismic-induced ore deposits

Regions that are or were seismic active and with the presence of fault zones might indicate there are ore deposits. A case study in Nevada, US by Howald (2015) studied how seismic-induced fluids accumulate mineral deposits, namely sinter and gold, along spaces provided by a fault zone. Two separate seismic events were identified and dated by oxygen isotopic concentrations, followed by episodes of the upward hydrothermal fluid migrations through permeable normal fault zone.[16] Mineralization started to take place when these hot silica-rich hydrothermal fluids met the cool meteoric water infiltrated along the fault zone until the convective flow system was shut down.[16] In order to deposit minerals, seismic events that bring hydrothermal fluids are not the only dominant factor, the permeability of the fault zone also has to be sufficient for permitting fluid flows.[16]

Another example taken from Sheldon (2005) also shows that development of fault zone, in this case by strike-slip faulting, facilitates mineralization. Sudden dilation happened along with strike-slip events increases the porosity and permeability along the fault zone.[31] Larger displacement will lead to greater increase in porosity.[31] If the faulting event cut through a sealing unit which seals a confined aquifer of over-pressured fluids, the fluids can rise through the fault zone.[31] Then mineralization will take place along the fault zone by pressure solution,[31] reducing the porosity of the fault zone. The fluid flow channel along the fault zone will be shut down when the pores are almost occupied by newly precipitated ore minerals.[31] Multiple seismic events have to be occurred to form these economic ore deposit with vein structure.[31]

See also

References

  1. 1.00 1.01 1.02 1.03 1.04 1.05 1.06 1.07 1.08 1.09 1.10 1.11 1.12 1.13 1.14 1.15 1.16 1.17 1.18 1.19 1.20 1.21 1.22 1.23 1.24 1.25 1.26 1.27 1.28 1.29 1.30 1.31 1.32 1.33 1.34 1.35 1.36 1.37 1.38 1.39 1.40 Bense, V.F.; Gleeson, T.; Loveless, S.E.; Bour, O.; Scibek, J. (2013). "Fault zone hydrogeology" (in en). Earth-Science Reviews 127: 171–192. doi:10.1016/j.earscirev.2013.09.008. Bibcode2013ESRv..127..171B. https://linkinghub.elsevier.com/retrieve/pii/S0012825213001657. 
  2. 2.00 2.01 2.02 2.03 2.04 2.05 2.06 2.07 2.08 2.09 2.10 2.11 2.12 2.13 2.14 Farrell, N.J.C.; Healy, D. (2017). "Anisotropic pore fabrics in faulted porous sandstones". Journal of Structural Geology 104: 125–141. doi:10.1016/j.jsg.2017.09.010. ISSN 0191-8141. Bibcode2017JSG...104..125F. 
  3. 3.0 3.1 3.2 3.3 3.4 3.5 3.6 Hadley, Daniel R.; Abrams, Daniel B.; Roadcap, George S. (2020). "Modeling a Large‐Scale Historic Aquifer Test: Insight into the Hydrogeology of a Regional Fault Zone" (in en). Groundwater 58 (3): 453–463. doi:10.1111/gwat.12922. ISSN 0017-467X. PMID 31290141. https://onlinelibrary.wiley.com/doi/abs/10.1111/gwat.12922. 
  4. 4.0 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 Scibek, J.; Gleeson, T.; McKenzie, J. M. (2016). "The biases and trends in fault zone hydrogeology conceptual models: global compilation and categorical data analysis" (in en). Geofluids 16 (4): 782–798. doi:10.1111/gfl.12188. Bibcode2016Gflui..16..782S. http://doi.wiley.com/10.1111/gfl.12188. 
  5. 5.0 5.1 Lin, Aiming; Yamashita, Kazuhiko (2013-12-01). "Spatial variations in damage zone width along strike-slip faults: An example from active faults in southwest Japan" (in en). Journal of Structural Geology 57: 1–15. doi:10.1016/j.jsg.2013.10.006. ISSN 0191-8141. Bibcode2013JSG....57....1L. http://www.sciencedirect.com/science/article/pii/S0191814113001776. 
  6. 6.0 6.1 6.2 6.3 6.4 6.5 Bu, F.; Xu, T.; Wang, F.; Yang, Z.; Tian, H. (2016). "Influence of highly permeable faults within a low-porosity and low-permeability reservoir on migration and storage of injected CO 2" (in en). Geofluids 16 (4): 769–781. doi:10.1111/gfl.12185. 
  7. 7.0 7.1 Achtziger-Zupančič, P.; Loew, S.; Hiller, A.; Mariethoz, G. (2016). "3D fluid flow in fault zones of crystalline basement rocks (Poehla-Tellerhaeuser Ore Field, Ore Mountains, Germany)" (in en). Geofluids 16 (4): 688–710. doi:10.1111/gfl.12192. 
  8. R.P. Dickerson (2000-10-19). Hydrologic Characteristics of Faults at Yucca Mountain, Nevada. doi:10.2172/860273. 
  9. Aydin, Atilla (2000-08-01). "Fractures, faults, and hydrocarbon entrapment, migration and flow" (in en). Marine and Petroleum Geology 17 (7): 797–814. doi:10.1016/S0264-8172(00)00020-9. ISSN 0264-8172. Bibcode2000MarPG..17..797A. http://www.sciencedirect.com/science/article/pii/S0264817200000209. 
  10. Faults and subsurface fluid flow in the shallow crust. Haneberg, William C.. Washington, DC: American Geophysical Union. 1999. ISBN 0-87590-096-8. OCLC 42061057. https://www.worldcat.org/oclc/42061057. 
  11. 11.0 11.1 11.2 11.3 11.4 Ingebritsen, S. E.; Manga, Michael (2019). "Earthquake Hydrogeology" (in en). Water Resources Research 55 (7): 5212–5216. doi:10.1029/2019WR025341. ISSN 1944-7973. Bibcode2019WRR....55.5212I. 
  12. Torabi, A.; Johannessen, M. U.; Ellingsen, T. S. S. (2019-06-02). "Fault Core Thickness: Insights from Siliciclastic and Carbonate Rocks" (in en). Geofluids 2019: 1–24. doi:10.1155/2019/2918673. 
  13. Shukla, Matsyendra Kumar; Sharma, Anupam (2018-06-01). "A brief review on breccia: it's [sic] contrasting origin and diagnostic signatures" (in en). Solid Earth Sciences 3 (2): 50–59. doi:10.1016/j.sesci.2018.03.001. ISSN 2451-912X. Bibcode2018SolES...3...50S. 
  14. 14.0 14.1 14.2 "Igneous Rock-forming Minerals", Earth Materials 2nd Edition (Cambridge University Press): pp. 159–196, 2016-12-15, doi:10.1017/9781316652909.009, ISBN 978-1-107-15540-4, http://dx.doi.org/10.1017/9781316652909.009, retrieved 2020-10-03 
  15. Blenkinsop, Tom G. (2000). Deformation microstructures and mechanisms in minerals and rocks. Dordrecht: Kluwer Academic Publishers. ISBN 978-0-412-73480-9. OCLC 70771466. https://www.worldcat.org/oclc/70771466. 
  16. 16.0 16.1 16.2 16.3 16.4 16.5 16.6 Howald, T.; Person, M.; Campbell, A.; Lueth, V.; Hofstra, A.; Sweetkind, D.; Gable, C. W.; Banerjee, A. et al. (2015). "Evidence for long timescale (>10 3 years) changes in hydrothermal activity induced by seismic events" (in en). Geofluids 15 (1–2): 252–268. doi:10.1111/gfl.12113. Bibcode2015Gflui..15..252H. 
  17. Shi, Zheming; Wang, Guangcai (2014). "Hydrological response to multiple large distant earthquakes in the Mile well, China" (in en). Journal of Geophysical Research: Earth Surface 119 (11): 2448–2459. doi:10.1002/2014JF003184. ISSN 2169-9011. Bibcode2014JGRF..119.2448S. https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1002/2014JF003184. 
  18. Wang, Chi-Yuen; Liao, Xin; Wang, Lee-Ping; Wang, Chung-Ho; Manga, Michael (2016). "Large earthquakes create vertical permeability by breaching aquitards" (in en). Water Resources Research 52 (8): 5923–5937. doi:10.1002/2016WR018893. ISSN 1944-7973. Bibcode2016WRR....52.5923W. 
  19. Shi, Zheming; Zhang, Shouchuan; Yan, Rui; Wang, Guangcai (2018-02-16). "Fault Zone Permeability Decrease Following Large Earthquakes in a Hydrothermal System" (in en). Geophysical Research Letters 45 (3): 1387–1394. doi:10.1002/2017GL075821. ISSN 0094-8276. Bibcode2018GeoRL..45.1387S. https://onlinelibrary.wiley.com/doi/abs/10.1002/2017GL075821. 
  20. "Does the production of oil and gas from shales cause earthquakes? If so, how are the earthquakes related to these operations? | U.S. Geological Survey". https://www.usgs.gov/faqs/does-production-oil-and-gas-shales-cause-earthquakes-if-so-how-are-earthquakes-related-these. 
  21. 21.0 21.1 21.2 21.3 21.4 21.5 21.6 21.7 21.8 Raduha, S.; Butler, D.; Mozley, P. S.; Person, M.; Evans, J.; Heath, J. E.; Dewers, T. A.; Stauffer, P. H. et al. (2016). "Potential seal bypass and caprock storage produced by deformation-band-to-opening-mode-fracture transition at the reservoir/caprock interface" (in en). Geofluids 16 (4): 752–768. doi:10.1111/gfl.12177. 
  22. 22.0 22.1 22.2 Ulmer-Scholle, Dana S.; Scholle, Peter A.; Schieber, Juergen; Raine, Robert J. (2015). A Color Guide to the Petrography of Sandstones, Siltstones, Shales and Associated Rocks. American Association of Petroleum Geologists. doi:10.1306/m1091304. ISBN 978-0-89181-389-7. http://dx.doi.org/10.1306/m1091304. 
  23. 23.0 23.1 Wang, Boyang; Zhang, Qiusheng; Wang, Gang; Liu, Miao (2018-08-23). "Analysis on the difference of material composition and reservoir space of Mesoproterozoic dark marine shale in the Yanshan area". Energy Exploration & Exploitation 37 (1): 332–354. doi:10.1177/0144598718797101. ISSN 0144-5987. 
  24. Chen, Xiu; Qu, Xiyu; Xu, Shouyu; Wang, Weiming; Li, Shunming; He, Hui; Liu, Yang (2020-07-01). "Dissolution pores in shale and their influence on reservoir quality in Damintun Depression, Bohai Bay Basin, East China: Insights from SEM images, N2 adsorption and fluid-rock interaction experiments" (in en). Marine and Petroleum Geology 117: 104394. doi:10.1016/j.marpetgeo.2020.104394. ISSN 0264-8172. Bibcode2020MarPG.11704394C. http://www.sciencedirect.com/science/article/pii/S026481722030177X. 
  25. 25.0 25.1 Musumeci, G.; Mazzarini, F.; Cruden, A. R. (2015). "The Zuccale Fault, Elba Island, Italy: A new perspective from fault architecture: THE BRITTLE ZUCCALE FAULT" (in en). Tectonics 34 (6): 1195–1218. doi:10.1002/2014TC003809. http://doi.wiley.com/10.1002/2014TC003809. 
  26. 26.0 26.1 Collettini, C.; Holdsworth, R.E. (2004). "Fault zone weakening and character of slip along low-angle normal faults: insights from the Zuccale fault, Elba, Italy" (in en). Journal of the Geological Society 161 (6): 1039–1051. doi:10.1144/0016-764903-179. ISSN 0016-7649. Bibcode2004JGSoc.161.1039C. http://jgs.lyellcollection.org/lookup/doi/10.1144/0016-764903-179. 
  27. 27.0 27.1 27.2 Lien, Tzuyi; Cheng, Ching-Chung; Hwang, Cheinway; Crossley, David (2014). "Assessing active faulting by hydrogeological modeling and superconducting gravimetry: A case study for Hsinchu Fault, Taiwan: Assess fault by hydrology and gravity" (in en). Journal of Geophysical Research: Solid Earth 119 (9): 7319–7335. doi:10.1002/2014JB011285. 
  28. 28.0 28.1 Elena, Fillimonova; Vasiliy, Lavrushin; Natalia, Kharitonova; Arslan, Sartykov; Elena, Maximova; Ekaterina, Baranovskaya; Anna, Korzun; Alexey, Maslov et al. (2019-12-11). "Hydrogeology and hydrogeochemistry of mineral sparkling groundwater within Essentuki area (Caucasian mineral water region)" (in en). Environmental Earth Sciences 79 (1): 15. doi:10.1007/s12665-019-8721-2. ISSN 1866-6299. https://doi.org/10.1007/s12665-019-8721-2. 
  29. Kim, J.; Ryan, P.; Klepeis, K.; Gleeson, T.; North, K.; Bean, J.; Davis, L.; Filoon, J. (2014). "Tectonic evolution of a Paleozoic thrust fault influences the hydrogeology of a fractured rock aquifer, northeastern Appalachian foreland" (in en). Geofluids 14 (3): 266–290. doi:10.1111/gfl.12076. Bibcode2014Gflui..14..266K. 
  30. Grizard, Pierre; Schmitt, Jean-Michel; Goblet, Patrick (2019-02-01). "Hydrogeology of an arid endorheic basin (Tsagaan Els, Dornogobi, Mongolia): field data and conceptualization, three-dimensional groundwater modeling, and water budget" (in en). Hydrogeology Journal 27 (1): 145–160. doi:10.1007/s10040-018-1868-1. ISSN 1435-0157. Bibcode2019HydJ...27..145G. https://doi.org/10.1007/s10040-018-1868-1. 
  31. 31.0 31.1 31.2 31.3 31.4 31.5 31.6 31.7 Sheldon, H. A.; Ord, A. (2005-10-20). "Evolution of porosity, permeability and fluid pressure in dilatant faults post-failure: implications for fluid flow and mineralization". Geofluids 5 (4): 272–288. doi:10.1111/j.1468-8123.2005.00120.x. ISSN 1468-8115.