Skip to main content

A Hydromechanical Model for Lower Crustal Fluid Flow

  • Chapter
  • First Online:

Part of the book series: Lecture Notes in Earth System Sciences ((LNESS))

Abstract

Metamorphic devolatilization generates fluids at, or near, lithostatic pressure. These fluids are ultimately expelled by compaction. It is doubtful that fluid generation and compaction operate on the same time scale at low metamorphic grade, even in rocks that are deforming by ductile mechanisms in response to tectonic stress. However, thermally-activated viscous compaction may dominate fluid flow patterns at moderate to high metamorphic grades. Compaction-driven fluid flow organizes into self-propagating domains of fluid-filled porosity that correspond to steady-state wave solutions of the governing equations. The effective rheology for compaction processes in heterogeneous rocks is dictated by the weakest lithology. Geological compaction literature invariably assumes linear viscous mechanisms; but lower crustal rocks may well be characterized by non-linear (power-law) viscous mechanisms. The steady-state solutions and scales derived here are general with respect to the dependence of the viscous rheology on effective pressure. These solutions are exploited to predict the geometry and properties of the waves as a function of rock rheology and the rate of metamorphic fluid production. In the viscous limit, wavelength is controlled by a hydrodynamic length scale δ, which varies inversely with temperature, and/or the rheological length scale for thermal activation of viscous deformation l A , which is on the order of a kilometer. At high temperature, such that δ < l A , waves are spherical. With falling temperature, as δ → l A , waves flatten to sill-like structures. If the fluid overpressures associated with viscous wave propagation reach the conditions for plastic failure, then compaction induces channelized fluid flow. The channeling is caused by vertically elongated porosity waves that nucleate with characteristic spacing δ. Because δ increases with falling temperature, this mechanism is amplified towards the surface. Porosity wave passage is associated with pressure anomalies that generate an oscillatory lateral component to the fluid flux that is comparable to the vertical component. As the vertical component may be orders of magnitude greater than time-averaged metamorphic fluxes, porosity waves are a potentially important agent for metasomatism. The time and spatial scales of these mechanisms depend on the initial state that is perturbed by the metamorphic process. Average fluxes place an upper limit on the spatial scale and a lower limit on the time scale, but the scales are otherwise unbounded. Thus, inversion of natural fluid flow patterns offers the greatest hope for constraining the compaction scales. Porosity waves are a self-localizing mechanism for deformation and fluid flow. In nature these mechanisms are superimposed on patterns induced by far-field stress and pre-existing heterogeneities.

This is a preview of subscription content, log in via an institution.

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   179.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   229.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD   249.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

References

  • Ague JJ, Baxter EF (2007) Brief thermal pulses during mountain building recorded by Sr diffusion in apatite and multicomponent diffusion in garnet. Earth Planet Sci Lett 261:500–516. doi:10.1016/j.epsl.2007.07.017

    Google Scholar 

  • Ague JJ, Park J, Rye DM (1998) Regional metamorphic dehydration and seismic hazard. Geophys Res Lett 25:4221–4224

    Google Scholar 

  • Aharonov E, Spiegelman M, Kelemen P (1997) Three-dimensional flow and reaction in porous media: implications for the earth’s mantle and sedimentary basins. J Geophys Res 102:14821–14833

    Google Scholar 

  • Ashby MF (1988) The modeling of hot isostatic pressing. In: Garvare T (ed) Proceedings hip: hot isostatic pressing – theories and applications. Centek, Lulea, pp 29–40

    Google Scholar 

  • Athy LF (1930) Density, porosity and compaction of sedimentary rocks. Bull Am Assoc Pet Geol 14:1–24

    Google Scholar 

  • Austrheim H (1987) Eclogitization of lower crustal granulites by fluid migration through shear zones. Earth Planet Sci Lett 81:221–232. doi:10.1016/0012-821X(87)90158-0

    Google Scholar 

  • Barcilon V, Lovera OM (1989) Solitary waves in magma dynamics. J Fluid Mech 204:121–133

    Google Scholar 

  • Barcilon V, Richter FM (1986) Nonlinear-waves in compacting media. J Fluid Mech 164:429–448

    Google Scholar 

  • Bercovici D, Ricard Y, Schubert G (2001) A two-phase model for compaction and damage 1. General theory. J Geophys Res 106:8887–8906

    Google Scholar 

  • Bouilhol P, Burg JP, Bodinier JL, Schmidt MW, Dawood H, Hussain S (2009) Magma and fluid percolation in arc to forearc mantle: evidence from Sapat (Kohistan, Northern Pakistan). Lithos 107:17–37. doi:10.1016/j.lithos.2008.07.004

    Google Scholar 

  • Bouilhol P, Connolly JAD, Burg JP (2011) Geological evidence and modeling of melt migration by porosity waves in the sub-arc mantle of Kohistan (Pakistan). Geology 39:1091–1094. doi:10.1130/G32219.1

    Google Scholar 

  • Cartwright I, Buick IS (1999) The flow of surface-derived fluids through alice springs age middle-crustal ductile shear zones, Reynolds Range, Central Australia. J Metamorph Geol 17:397–414

    Google Scholar 

  • Cheadle MJ, Elliott MT, McKenzie D (2004) Percolation threshold and permeability of crystallizing igneous rocks: the importance of textural equilibrium. Geology 32:757–760. doi:10.1130/g20495.1

    Google Scholar 

  • Connolly JAD (1997) Devolatilization-generated fluid pressure and deformation-propagated fluid flow during prograde regional metamorphism. J Geophys Res 102:18149–18173

    Google Scholar 

  • Connolly JAD (2010) The mechanics of metamorphic fluid expulsion. Elements 6:165–172. doi:10.2113/gselements.6.3.165

    Google Scholar 

  • Connolly JAD, Podladchikov YY (1998) Compaction-driven fluid flow in viscoelastic rock. Geodinamica Acta 11:55–84

    Google Scholar 

  • Connolly JAD, Podladchikov YY (2000) Temperature-dependent viscoelastic compaction and compartmentalization in sedimentary basins. Tectonophysics 324:137–168

    Google Scholar 

  • Connolly JAD, Podladchikov YY (2004) Fluid flow in compressive tectonic settings: implications for mid-crustal seismic reflectors and downward fluid migration. J Geophys Res 109:B04201. doi:10.1029/2003JB002822

    Google Scholar 

  • Connolly JAD, Podladchikov YY (2007) Decompaction weakening and channeling instability in ductile porous media: implications for asthenospheric melt segregation. J Geophys Res 112:B10205. doi:10.1029/2005jb004213

    Google Scholar 

  • Connolly JAD, Thompson AB (1989) Fluid and enthalpy production during regional metamorphism. Contrib Mineral Petrol 102:346–366

    Google Scholar 

  • Connolly JAD, Schmidt MW, Solferino G, Bagdassarov N (2009) Permeability of asthenospheric mantle and melt extraction rates at mid-ocean ridges. Nature 462:209–212. doi:10.1038/nature08517

    Google Scholar 

  • Cox SF (2005) Coupling between deformation, fluid pressures and fluid flow in ore-producing hydrothermal environments. In: Economic geology 100th anniversary volume, Denver, pp 39–75

    Google Scholar 

  • Cox SF, Ruming K (2004) The St Ives mesothermal gold system, Western Australia – a case of golden aftershocks? J Struct Geol 26:1109–1125. doi:10.1016/j.jsg.2003.11.025

    Google Scholar 

  • Daines MJ, Kohlstedt DL (1994) The transition from porous to channelized flow due to melt/rock reaction during melt migration. Geophys Res Lett 21:145–148

    Google Scholar 

  • Dewey JF (2005) Orogeny can be very short. Proc Natl Acad Sci USA 102:15286–15293. doi:10.1073/pnas.0505516102

    Google Scholar 

  • England PC, Thompson AB (1984) Pressure temperature time paths of regional metamorphism. 1. Heat-transfer during the evolution of regions of thickened continental-crust. J Petrol 25:894–928

    Google Scholar 

  • Etheridge MA, Wall VJ, Cox SF, Vernon RH (1984) High fluid pressures during regional metamorphism and deformation – implications for mass-transport and deformation mechanisms. J Geophys Res 89:4344–4358

    Google Scholar 

  • Ferry JM (1994) Overview of the petrologic record of fluid-flow during regional metamorphism in Northern New-England. Am J Sci 294:905–988

    Google Scholar 

  • Ferry JM, Gerdes ML (1998) Chemically reactive fluid flow during metamorphism. Annu Rev Earth Planet Sci 26:255–287

    Google Scholar 

  • Fulton PM, Saffer DM, Bekins BA (2009) A critical evaluation of crustal dehydration as the cause of an overpressured and weak San Andreas Fault. Earth Planet Sci Lett 284:447–454. doi:10.1016/j.epsl.2009.05.009

    Google Scholar 

  • Gavrilenko P, Gueguen Y (1993) Fluid overpressures and pressure solution in the crust. Tectonophysics 21:91–110

    Google Scholar 

  • Gleeson SA, Yardley BWD, Boyce AJ, Fallick AE, Munz LA (2000) From basin to basement: the movement of surface fluids into the crust. J Geochem Explor 69:527–531

    Google Scholar 

  • Gleeson SA, Yardley BWD, Munz IA, Boyce AJ (2003) Infiltration of basinal fluids into high-grade basement, South Norway: sources and behaviour of waters and brines. Geofluids 3:33–48

    Google Scholar 

  • Gold T, Soter S (1985) Fluid ascent through the solid lithosphere and its relation to earthquakes. Pure Appl Geophys 122:492–530

    Google Scholar 

  • Goldschmidt VM (1954) Geochemistry. Clarendon, Oxford, p 730

    Google Scholar 

  • Graham CM, Valley JW, Eiler JM, Wada H (1998) Timescales and mechanisms of fluid infiltration in a marble: an ion microprobe study. Contrib Mineral Petrol 132:371–389

    Google Scholar 

  • Gratier JP, Favreau P, Renard F (2003) Modeling fluid transfer along California faults when integrating pressure solution crack sealing and compaction processes. J Geophys Res 108. doi:10.1029/2001jb000380

  • Gueguen Y, Palciauskas VV (1994) Introduction to the physics of rocks. Princeton University Press, Princeton, p 194

    Google Scholar 

  • Gueguen Y, Dormieux L, Bouteca M (2004) Fundamentals of poromechanics. In: Gueguen Y, Bouteca M (eds) Mechanics of fluid-saturated rocks. Elsevier Academic, Burlington, pp 55–79

    Google Scholar 

  • Hammer PTC, Clowes RM (1996) Seismic reflection investigations of the mount cayley bright spot: a midcrustal reflector beneath the Coast Mountains, British Columbia. J Geophys Res 101:20119–20131

    Google Scholar 

  • Hanson RB (1997) Hydrodynamics of regional metamorphism doe to continental collision. Econ Geol Bull Soc Econ Geol 92:880–891

    Google Scholar 

  • Hetenyi G, Cattin R, Brunet F, Bollinger L, Vergne J, Nabelek J, Diament M (2007) Density distribution of the india plate beneath the Tibetan Plateau: geophysical and petrological constraints on the kinetics of lower-crustal eclogitization. Earth Planet Sci Lett 264:226–244. doi:10.1016/j.epst.2007.09.036

    Google Scholar 

  • Hill R (1950) The mathematical theory of plasticity. Clarendon, Oxford, p 356

    Google Scholar 

  • Holness MB, Siklos STC (2000) The rates and extent of textural equilibration in high-temperature fluid-bearing systems. Chem Geol 162:137–153. doi:10.1016/s0009-2541(99)00124-2

    Google Scholar 

  • Holtzman BK, Kohlstedt DL (2007) Stress-driven melt segregation and strain partitioning in partially molten rocks: effects of stress and strain. J Petrol 48:2379–2406. doi:10.1093/petrology/egm065

    Google Scholar 

  • Huenges E, Erzinger J, Kuck J, Engeser B, Kessels W (1997) The permeable crust: geohydraulic properties down to 9101 m depth. J Geophys Res 102:18255–18265

    Google Scholar 

  • Hunt JM (1990) Generation and migration of petroleum from abnormally pressured fluid compartments. Am Assoc Pet Geol 74:1–12

    Google Scholar 

  • Ingebritsen SE, Manning CE (1999) Geological implications of a permeability-depth curve for the continental crust. Geology 27:1107–1110

    Google Scholar 

  • Ingebritsen SE, Manning CE (2010) Permeability of the continental crust: dynamic variations inferred from seismicity and metamorphism. Geofluids 10:193–205. doi:10.1111/j.1468-8123.2010.00278.x

    Google Scholar 

  • Ingebritsen SE, Sanford WE, Neuzil C (2006) Groundwater in geologic processes. Cambridge University Press, Cambridge

    Google Scholar 

  • Jagoutz O, Muntener O, Burg JP, Ulmer P, Jagoutz E (2006) Lower continental crust formation through focused flow in km-scale melt conduits: the zoned ultramafic bodies of the Chilas complex in the Kohistan Island Arc (NW Pakistan). Earth Planet Sci Lett 242:320–342

    Google Scholar 

  • Jiracek GR, Gonzalez VM, Caldwell TG, Wannamaker PE, Kilb D (2007) Seismogenic, electrically conductive, and fluid zones at continental plate boundaries in New Zealand, Himalaya, and California-USA. Geophys Monogr Ser 175:347–369. doi:10.1029/175GM18

    Google Scholar 

  • Kerschhofer L, Dupas C, Liu M, Sharp TG, Durham WB, Rubie DC (1998) Polymorphic transformations between olivine, wadsleyite and ringwoodite: mechanisms of intracrystalline nucleation and the role of elastic strain. Mineral Mag 62:617–638

    Google Scholar 

  • Kohlstedt DL, Evans B, Mackwell SJ (1995) Strength of the lithosphere – constraints imposed by laboratory experiments. J Geophys Res 100:17587–17602

    Google Scholar 

  • Liang Y, Schiemenz A, Hesse MA, Parmentier EM, Hesthaven JS (2010) High-porosity channels for melt migration in the mantle: top is the dunite and bottom is the harzburgite and lherzolite. Geophys Res Lett 37:L15306. doi:10.1029/2010gl044162

    Google Scholar 

  • Liotta D, Ranalli G (1999) Correlation between seismic reflectivity and rheology in extended lithosphere: southern Tuscany, Inner Northern Apennines, Italy. Tectonophysics 315:109–122

    Google Scholar 

  • Makovsky Y, Klemperer SL (1999) Measuring the seismic properties of tibetan bright spats: evidence for free aqueous fluids in the Tibetan middle crust. J Geophys Res 104:10795–10825

    Google Scholar 

  • Mancktelow NS (2008) Tectonic pressure: theoretical concepts and modelled examples. Lithos 103:149–177. doi:10.1016/j.lithos.2007.09.013

    Google Scholar 

  • Manning CE, Ingebritsen SE (1999) Permeability of the continental crust: implications of geothermal data and metamorphic systems. Rev Geophys 37:127–150

    Google Scholar 

  • McCaig AM, Wickham SM, Taylor HP (1990) Deep fluid circulation in alpine shear zones, Pyrenees, France – field and oxygen isotope studies. Contrib Mineral Petrol 106:41–60

    Google Scholar 

  • McCuaig TC, Kerrich R (1998) P-T-t-deformation-fluid characteristics of lode gold deposits: evidence from alteration systematics. Ore Geol Rev 12:381–453

    Google Scholar 

  • McKenzie D (1984) The generation and compaction of partially molten rock. J Petrol 2:713–765

    Google Scholar 

  • Milke R, Abart R, Kunze K, Koch-Muller M, Schmid D, Ulmer P (2009) Matrix rheology effects on reaction rim growth i: evidence from orthopyroxene rim growth experiments. J Metamorph Geol 27:71–82. doi:10.1111/j.1525-1314.2008.00804.x

    Google Scholar 

  • Miller SA, Collettini C, Chiaraluce L, Cocco M, Barchi M, Kaus BJP (2004) Aftershocks driven by a high-pressure CO2 source at depth. Nature 427:724–727. doi:10.1038/nature02251

    Google Scholar 

  • Mosenfelder JL, Connolly JAD, Rubie DC, Liu M (2000) Strength of (Mg, Fe)2SiO4 wadsleyite determined by relaxation of transformation stress. Phys Earth Planet Inter 120:63–78

    Google Scholar 

  • Munz IA, Yardley BWD, Gleeson SA (2002) Petroleum infiltration of high-grade basement, South Norway: pressure-temperature-time-composition (P-T-t-X) constraints. Geofluids 2:41–53

    Google Scholar 

  • Nabelek PI (2009) Numerical simulation of kinetically-controlled calc-silicate reactions and fluid flow with transient permeability around crystallizing plutons. Am J Sci 309:517–548. doi:10.2475/07.2009.01

    Google Scholar 

  • Nakashima Y (1995) Transport model of buoyant metamorphic fluid by hydrofracturing in leaky rock. J Metamorph Geol 13:727–736

    Google Scholar 

  • Neuzil CE (1994) How permeable are clays and shales? Water Resour Res 30:145–150

    Google Scholar 

  • Norton D, Knapp R (1977) Transport phenomena in hydrothermal systems – nature of porosity. Am J Sci 277:913–936

    Google Scholar 

  • Okamoto A, Tsuchiya N (2009) Velocity of vertical fluid ascent within vein-forming fractures. Geology 37:563–566. doi:10.1130/g25680a.1

    Google Scholar 

  • Oliver GJH, Chen F, Buchwaldt R, Hegner E (2000) Fast tectonometamorphism and exhumation in the type area of the Barrovian and Buchan zones. Geology 28:459–462

    Google Scholar 

  • Ozel O, Iwasaki T, Moriya T, Sakai S, Maeda T, Piao C, Yoshii T, Tsukada S, Ito A, Suzuki M, Yamazaki A, Miyamachi H (1999) Crustal structure of Central Japan and its petrological implications. Geophys J Int 138:257–274

    Google Scholar 

  • Padron-Navarta JA, Tommasi A, Garrido CJ, Sanchez-Vizcaino VL, Gomez-Pugnaire MT, Jabaloy A, Vauchez A (2010) Fluid transfer into the wedge controlled by high-pressure hydrofracturing in the cold top-slab mantle. Earth Planet Sci Lett 297:271–286. doi:10.1016/j.epsl.2010.06.029

    Google Scholar 

  • Padron-Navarta JA, Sanchez-Vizcaino VL, Garrido CJ, Gomez-Pugnaire MT (2011) Metamorphic record of high-pressure dehydration of antigorite serpentinite to chlorite harzburgite in a subduction setting (Cerro del Almirez, Nevado-Filabride Complex, Southern Spain). J Petrol 52:2047–2078. doi:10.1093/petrology/egr039

    Google Scholar 

  • Paterson MS, Luan FC (1990) Quartzite rheology under geological conditions. In: Knipe RJ, Rutter EH (eds) Deformation mechanisms, rheology and tectonics, vol 54. The Geological Society, London, pp 299–307

    Google Scholar 

  • Peng ZG, Vidale JE, Creager KC, Rubinstein JL, Gomberg J, Bodin P (2008) Strong tremor near Parkfield, CA, excited by the 2002 Denali fault earthquake. Geophys Res Lett 35. doi:10.1029/2008gl036080

  • Petrini K, Podladchikov Y (2000) Lithospheric pressure-depth relationship in compressive regions of thickened crust. J Metamorph Geol 18:67–78

    Google Scholar 

  • Plank T, Langmuir CH (1998) The chemical composition of subducting sediment and its consequences for the crust and mantle. Chem Geol 145:325–394

    Google Scholar 

  • Powley DE (1990) Pressures and hydrogeology in petroleum basins. Earth Sci Rev 29:215–226

    Google Scholar 

  • Price JD, Wark DA, Watson EB, Smith AM (2006) Grain-scale permeabilities of faceted polycrystalline aggregates. Geofluids 6:302–318. doi:10.1111/j.1468-8123.2006.00149.x

    Google Scholar 

  • Ranalli G (1995) Rheology of the earth. Springer, New York

    Google Scholar 

  • Ranalli G, Rybach L (2005) Heat flow, heat transfer and lithosphere rheology in geothermal areas: features and examples. J Volcanol Geotherm Res 148:3–19. doi:10.1016/j.jvolgeores.2005.04.010

    Google Scholar 

  • Read CM, Cartwright I (2000) Meteoric fluid infiltration in the middle crust during shearing: examples from the Arunta Inlier, Central Australia. J Geochem Explor 69:333–337

    Google Scholar 

  • Rice JR (1992) Fault stress states, pore pressure distributions, and the weakness of the San Andreas Fault. In: Evans B, Wong T-F (eds) Fault mechanics and transport properties of rocks. Academic, New York, pp 475–503

    Google Scholar 

  • Richter FM, McKenzie D (1984) Dynamical models for melt segregation from a deformable rock matrix. J Geol 92:729–740

    Google Scholar 

  • Rozhko AY, Podladchikov YY, Renard F (2007) Failure patterns caused by localized rise in pore-fluid overpressure and effective strength of rocks. Geophys Res Lett 34:L22304. doi:10.1029/2007gl031696

    Google Scholar 

  • Rubie DC, Thompson AB (1985) Kinetics of metamorphic reactions. In: Thompson AB, Rubie DC (eds) Metamorphic reactions kinetics, textures, and deformation, vol 4. Springer, New York, pp 291

    Google Scholar 

  • Rubin RM (1995) Propagation of magma-filled cracks. Annu Rev Earth Planet Sci 23:287–336

    Google Scholar 

  • Rubinstein JL, Vidale JE, Gomberg J, Bodin P, Creager KC, Malone SD (2007) Non-volcanic tremor driven by large transient shear stresses. Nature 448:579–582. doi:10.1038/nature06017

    Google Scholar 

  • Rutter EH (1983) Pressure solution in nature, theory and experiment. J Geol Soc Lond 140:725–740

    Google Scholar 

  • Scarpa R, Amoruso A, Crescentini L, Fischione C, Formisano LA, La Rocca M, Tronca F (2008) Slow earthquakes and low frequency tremor along the Apennines, Italy. Ann Geophys 51:527–538

    Google Scholar 

  • Schmalholz SM, Podladchikov Y (1999) Buckling versus folding: importance of viscoelasticity. Geophys Res Lett 26:2641–2644

    Google Scholar 

  • Scholz CH (1988) The brittle-plastic transition and the depth of seismic faulting. Geol Rundsch 77:319–328

    Google Scholar 

  • Scott DR (1988) The competition between percolation and circulation in a deformable porous-medium. J Geophys Res 93:6451–6462

    Google Scholar 

  • Scott DR, Stevenson DJ (1986) Magma ascent by porous flow. J Geophys Res 91:9283–9296

    Google Scholar 

  • Scott DR, Stevenson DJ, Whitehead JA (1986) Observations of solitary waves in a viscously deformable pipe. Nature 319:759–761

    Google Scholar 

  • Shaw DM (1956) Geochemistry of pelite rocks, iii, major elements and general geochemistry. Geol Soc Am Bull 67:919–934

    Google Scholar 

  • Shimizu I (1995) Kinetics of pressure solution creep in quartz; theoretical considerations. Tectonophysics 245:121–134

    Google Scholar 

  • Sibson RH (1986) Earthquakes and rock deformation in crustal fault zones. Annu Rev Earth Planet Sci 14:149–175

    Google Scholar 

  • Sibson RH (1992) Fault-valve behavior and the hydrostatic lithostatic fluid pressure interface. Earth Sci Rev 32:141–144

    Google Scholar 

  • Sibson RH (2000) Tectonic controls on maximum sustainable overpressure: fluid redistribution from stress transitions. J Geochem Explor 69:471–475

    Google Scholar 

  • Sibson RH (2004) Controls on maximum fluid overpressure defining conditions for mesozonal mineralisation. J Struct Geol 26:1127–1136. doi:10.1016/j.jsg.2003.11.003

    Google Scholar 

  • Sibson RH (2009) Rupturing in overpressured crust during compressional inversion-the case from NE Honshu, Japan. Tectonophysics 473:404–416. doi:10.1016/j.tecto.2009.03.016

    Google Scholar 

  • Simpson GDH (1998) Dehydration-related deformation during regional metamorphism, NW Sardinia, Italy. J Metamorph Geol 16:457–472

    Google Scholar 

  • Skelton ADL (1996) The timing and direction of metamorphic fluid flow in Vermont. Contrib Mineral Petrol 125:75–84

    Google Scholar 

  • Skelton ADL, Valley JW, Graham CM, Bickle MJ, Fallick AE (2000) The correlation of reaction and isotope fronts and the mechanism of metamorphic fluid flow. Contrib Mineral Petrol 138:364–375

    Google Scholar 

  • Spiegelman M (1993) Flow in deformable porous-media. 1. Simple analysis. J Fluid Mech 247:17–38

    Google Scholar 

  • Spiegelman M, Elliott T (1993) Consequences of melt transport for uranium series disequilibrium in young lavas. Earth Planet Sci Lett 118:1–20

    Google Scholar 

  • Spiegelman M, Kelemen PB, Aharonov E (2001) Causes and consequences of flow organization during melt transport: the reaction infiltration instability in compactible media. J Geophys Res 106:2061–2077

    Google Scholar 

  • Spiers CJ, Schutjens PMTM (1990) Densification of crystalline aggregates by diffusional creep. J Geolo Soc Spec Publ 54:215–227

    Google Scholar 

  • Stahle HJ, Raith M, Hoernes S, Delfs A (1987) Element mobility during incipient granulite formation at kabbaldurga, Southern India. J Petrol 28:803–834

    Google Scholar 

  • Staude S, Bons PD, Markl G (2009) Hydrothermal vein formation by extension-driven dewatering of the middle crust: an example from SW Germany. Earth Planet Sci Lett 286:387–395. doi:10.1016/j.epsl.2009.07.012

    Google Scholar 

  • Stern T, Kleffmann S, Okaya D, Scherwath M, Bannister S (2001) Low seismic-wave speeds and enhanced fluid pressure beneath the Southern Alps of New Zealand. Geology 29:679–682

    Google Scholar 

  • Stevenson D (1989) Spontaneous small-scale melt segregation in partial melts undergoing deformation. Geophys Res Lett 16:1067–1070

    Google Scholar 

  • Suetnova EI, Carbonell R, Smithson SB (1994) Bright seismic reflections and fluid movement by porous flow in the lower crust. Earth Planet Sci Lett 126:161–169. doi:10.1016/0012-821x(94)90248-8

    Google Scholar 

  • Sumita I, Yoshida S, Kumazawa M, Hamano Y (1996) A model for sedimentary compaction of a viscous medium and its application to inner-core growth. Geophys J Int 124:502–524

    Google Scholar 

  • Tenthorey E, Cox SF (2006) Cohesive strengthening of fault zones during the interseismic period: an experimental study. J Geophys Res 111:B09202. doi:10.1029/2005jb004122

    Google Scholar 

  • Thompson AB, Connolly JAD (1990) Metamorphic fluids and anomalous porosities in the lower crust. Tectonophysics 182:47–55

    Google Scholar 

  • Tumarkina E, Misra S, Burlini L, Connolly JAD (2011) An experimental study of the role of shear deformation on partial melting of a synthetic metapelite. Tectonophysics 503:92–99. doi:10.1016/j.tecto.2010.12.004

    Google Scholar 

  • Upton P, Koons PO, Chamberlain CP (1995) Penetration of deformation-driven meteoric water into ductile rocks: isotopic and model observations from the Southern Alps, New Zealand. N Z J Geol Geophys 38:535–543

    Google Scholar 

  • van Haren JLM, Ague JJ, Rye DM (1996) Oxygen isotope record of fluid infiltration and mass transfer during regional metamorphism of pelitic schist, Connecticut, USA. Geochim et Cosmochim Acta 60:3487–3504

    Google Scholar 

  • Vanyan LL, Gliko AO (1999) Seismic and electromagnetic evidence of dehydration as a free water source in the reactivated crust. Geophys J Int 137:159–162

    Google Scholar 

  • Vrijmoed JC, Podladchikov YY, Andersen TB, Hartz EH (2009) An alternative model for ultra-high pressure in the Svartberget Fe-Ti Garnet-Peridotite, Western Gneiss Region, Norway. Eur J Mineral 21:1119–1133. doi:10.1127/0935-1221/2009/0021-1985

    Google Scholar 

  • Walther JV, Orville PM (1982) Volatile production and transport in regional metamorphism. Contrib Mineral Petrol 79:252–257

    Google Scholar 

  • Wark DA, Watson EB (1998) Grain-scale permeabilities of texturally equilibrated, monomineralic rocks. Earth Planet Sci Lett 164:591–605

    Google Scholar 

  • Warren CJ, Smye AJ, Kelley SP, Sherlock SC (2011) Using white mica 40Ar/39Ar data as a tracer for fluid flow and permeability under high-p conditions: Tauern Window, Eastern Alps. J Metamorph Geol. doi:10.1111/j.1525-1314.2011.00956.x

  • Wickham SM, Peters MT, Fricke HC, Oneil JR (1993) Identification of magmatic and meteoric fluid sources and upward-moving and downward-moving infiltration fronts in a metamorphic core complex. Geology 21:81–84

    Google Scholar 

  • Wiggins C, Spiegelman M (1995) Magma migration and magmatic solitary waves in 3-d. Geophys Res Lett 22:1289–1292

    Google Scholar 

  • Wilkinson DS, Ashby MF (1975) Pressure sintering by power law creep. Acta Metall 23:1277–1285

    Google Scholar 

  • Wing BA, Ferry JM (2007) Magnitude and geometry of reactive fluid flow from direct inversion of spatial patterns of geochemical alteration. Am J Sci 307:793–832. doi:10.2475/05.2007.02

    Google Scholar 

  • Wong TF, Christian D, Menendez B (2004) Mechanical compaction. In: Gueguen Y, Bouteca M (eds) Mechanics of fluid-saturated rocks. Elsevier Academic, Burlington, pp 55–79

    Google Scholar 

  • Xiao XH, Evans B, Bernabe Y (2006) Permeability evolution during non-linear viscous creep of calcite rocks. Pure Appl Geophys 163:2071–2102. doi:10.1007/s00024-006-0115-1

    Google Scholar 

  • Yardley BWD (1983) Quartz veins and devolatilization during metamorphism. J Geol Soc 140:657–663

    Google Scholar 

  • Yardley B, Gleeson S, Bruce S, Banks D (2000) Origin of retrograde fluids in metamorphic rocks. J Geochem Explor 69:281–285

    Google Scholar 

  • Young ED, Rumble D (1993) The origin of correlated variations in insitu O18/O16 and elemental concentrations in metamorphic garnet from southeastern Vermont, USA. Geochim et Cosmochim Acta 57:2585–2597. doi:10.1016/0016-7037(93)90419-w

    Google Scholar 

  • Zharikov AV, Vitovtova VM, Shmonov VM, Grafchikov AA (2003) Permeability of the rocks from the Kola superdeep borehole at high temperature and pressure: implication to fluid dynamics in the continental crust. Tectonophysics 370:177–191. doi:10.1016/s0040-1951(03)00185-9

    Google Scholar 

  • Zhu W, David C, T-f W (1995) Network modeling of permeability evolution during cementation and hot isostatic pressing. J Geophys Res 100:15451–15464

    Google Scholar 

  • Zhu W, Evans B, Bernabe Y (1999) Densification and permeability reduction in hot-pressed calcite: a kinetic model. J Geophys Res 104:25501–25511

    Google Scholar 

  • Zoback MD, Townend J (2001) Implications of hydrostatic pore pressures and high crustal strength for the deformation of intraplate lithosphere. Tectonophysics 336:19–30

    Google Scholar 

Download references

Acknowledgements

This paper was improved by outstanding, although not necessarily laudatory, reviews by Stephen F. Cox, Yves Gueguen, Peter O. Koons and Peter I. Nabelek. I am grateful to Dan Harlov and Hakon Austrheim for their extraordinary patience and judicious editing. The work presented here was supported by Swiss National Science Foundation grant 200021_130411.

Author information

Authors and Affiliations

Authors

Appendix: Steady-State Porosity Waves in a Viscous Matrix

Appendix: Steady-State Porosity Waves in a Viscous Matrix

This appendix presents a steady-state wave solution for flow of an incompressible fluid through a viscous matrix composed of incompressible solid grains. Geological compaction literature invariably assumes Newtonian behavior for the viscous mechanism; however, lower crustal environments may well be characterized by power-law viscous rheology (e.g., Kohlstedt et al. 1995). Accordingly, the solution derived here is general with respect to the dependence of the viscous rheology on effective pressure. Aside from this modification, the mathematical formulation of the governing compaction equations is identical to that of Connolly and Podladchikov (2000, 2007).

Conservation of solid and fluid mass requires

$$ \frac{{\partial (1 - {\phi} )}}{{\partial t}} + \nabla \cdot \left( {\left( {1 - {\phi} } \right){{{\mathbf{v}}}_{\rm{s}}}} \right) = 0 $$
(14.35)

and

$$ \frac{{\partial {\phi} }}{{\partial t}} + \nabla \cdot \left( {{\phi} {{{\mathbf{v}}}_{\rm{f}}}} \right) = 0, $$
(14.36)

where subscripts f and s distinguish the velocities, v, of the fluid and matrix. From Darcy’s law, force balance between the matrix and fluid requires

$$ {\mathbf{q}} = {\phi} \left( {{{{\mathbf{v}}}_{\rm{f}}} - {{{\mathbf{v}}}_{\rm{s}}}} \right) = - \frac{k}{\mu }\left( {\nabla {{p}_{\rm{f}}} - {{{\rho} }_{\rm{f}}}{\hbox{g}}{{{\mathbf{u}}}_{\rm{z}}}} \right). $$
(14.37)

In one-dimensional compaction of a vertical column, mean stress is identical to the load

$$ \bar{\sigma } = \int\limits_0^z {\left( {\left( {1 - {\phi} } \right){{{\rho} }_{\rm{s}}} + {\phi} {{{\rho} }_{\rm{f}}}} \right) } {\hbox{g}}{{{\mathbf{u}}}_{\rm{z}}}{\hbox{d}}z. $$
(14.38)

Thus, in terms of effective pressure, \( {{p}_{\rm{e}}} = \bar{\sigma } - {{p}_{\rm{f}}}, \) Eq. 14.37 is

$$ {\phi} \left( {{{{\mathbf{v}}}_{\rm{f}}} - {{{\mathbf{v}}}_{\rm{s}}}} \right) = \frac{k}{\mu }\left( {\nabla {{p}_{\rm{e}}} - \left( {1 - {\phi} } \right)\Delta {\rho} {\hbox{g}}{{{\mathbf{u}}}_{\rm{z}}}} \right). $$
(14.39)

The divergence of the total volumetric flux of matter is the sum of Eqs. 14.35 and 14.36

$$ \nabla \cdot \left( {{{{\mathbf{v}}}_{\rm{s}}} + {\phi} \left( {{{{\mathbf{v}}}_{\rm{f}}} - {{{\mathbf{v}}}_{\rm{s}}}} \right)} \right) = 0, $$
(14.40)

and substituting Eq. 14.39 into Eq. 14.40

$$ \nabla \cdot \left( {{{{\mathbf{v}}}_{\rm{s}}} + \frac{k}{\mu }\left( {\nabla {{p}_{\rm{e}}} - \left( {1 - {\phi} } \right)\Delta {\rho} {\hbox{g}}{{{\mathbf{u}}}_{\rm{z}}}} \right)} \right) = 0. $$
(14.41)

Matrix rheology is introduced with Eq. 14.16 by observing that the divergence of the solid velocity is the dilational strain rate of the matrix

$$ \nabla \cdot {{{\mathbf{v}}}_{\rm{s}}} = \frac{{\phi} }{{1 - {\phi} }}{{\dot{\varepsilon }}_{\phi}} = - {c}_{\sigma}{{f}_{\phi}}A{{\left| {p_{\rm{e}}} \right|}^{{{{n}_{\sigma }} - 1}}}p_{\rm{e}}^{{}} $$
(14.42)

where \( {{f}_{\phi}} = {\phi} \left( {1 - {\phi} } \right)/{{\left( {1 - {{{\phi} }^{{1/{{n}_{\sigma }}}}}} \right)}^{{{{n}_{\sigma }}}}} \)(Wilkinson and Ashby 1975). As the functional form of Eq. 14.42 may vary depending on the magnitude of the porosity and the viscous mechanism (Ashby 1988), the subsequent analysis assumes f φ is an unspecified function of porosity.

To avoid the unnecessary complication associated with the use of vector notation for a one-dimensional problem, in the remainder of this analysis vector quantities are represented by signed scalars and the gradient and divergence operators are replaced by ∂/∂z. Supposing the existence of a steady state solution in which fluid explusion is accomplished by waves that propagate with unchanging form through a matrix with background porosity φ0 filled by fluid at zero effective pressure, then, in a reference frame that travels with the wave, integration of Eq. 14.35 gives the matrix velocity as

$$ {{v}_{\rm{s}}} = {{v}_{\infty }}\frac{{1 - {{{\phi} }_0}}}{{1 - {\phi} }} $$
(14.43)

where v is the solid velocity in the limits φ → φ0 and p e → 0, i.e., at infinite distance from the wave. After substitution of Eq. 14.43, the integrated form of Eq. 14.41 can be rearranged to

$$ \frac{{\partial {{p}_{\rm{e}}}}}{{\partial z}} = \left( {{{q}_{\rm{t}}} - {{v}_{\infty }}\frac{{1 - {{{\phi} }_0}}}{{1 - {\phi} }}} \right)\frac{\mu }{k} + \left( {1 - {\phi} } \right)\Delta \rho {\hbox{g}} $$
(14.44)

where q t = φv f + (1 − φ)v s is the constant, total, volumetric flux of matter through the column, which evaluates in the limit φ → φ0 and p e → 0 as

$$ {{q}_{\rm{t}}} = {{v}_{\infty }} - \left( {1 - {{{\phi} }_{{0}}}} \right)\frac{{{{k}_0}}}{\mu }\Delta \rho {\hbox{g}} $$
(14.45)

where k 0 is the permeability at φ0. Thus, Eq. 14.44 can be rewritten

$$ \frac{{\partial {{p}_{\rm{e}}}}}{{\partial z}} = \Delta {\rho} {\hbox{g}}\left( {1 - {\phi} - \left( {1 - {{{\phi} }_0}} \right)\frac{{{{k}_0}}}{k}} \right) - {{v}_{\infty }}\frac{\mu }{k}\frac{{{\phi} - {{{\phi} }_0}}}{{1 - {\phi} }}. $$
(14.46)

Likewise, substitution of Eq. 14.43 into Eq. 14.42 yields

$$ \frac{{\partial {\phi} }}{{\partial z}} = - \frac{{{{{\left( {1 - {\phi} } \right)}}^2}}}{{1 - {{{\phi} }_0}}}{{f}_{\phi}}\frac{{{{c}_{\sigma }}A{{{\left| {p_{\rm{e}}} \right|}}^{{{{n}_{\sigma }} - 1}}}p_{\rm{e}}^{{}}}}{{{{v}_{\infty }}}} $$
(14.47)

If permeability is an, as yet unspecified, function of porosity, then Eqs. 14.46 and 14.47 form a closed system of two ordinary differential equations in two unknown functions, φ and p e. As v is the solid velocity at infinite distance from a steady-state wave, if the reference frame is changed to that of the unperturbed matrix, the phase velocity of the wave is v φ = −v .

For notational simplicity Eqs. 14.46 and 14.47 are now rewritten as

$$ \frac{{\partial {{p}_{\rm{e}}}}}{{\partial z}} = {{f}_1} $$
(14.48)
$$ \frac{{\partial {\phi} }}{{\partial z}} = {{f}_2}\frac{{{{c}_{\sigma }}A}}{{{{v}_{\phi}}}}{{\left| {p_{\rm{e}}} \right|}^{{{{n}_{\sigma }} - 1}}}p_{\rm{e}}^{{}} $$
(14.49)

where f 1 is the dependence of Eq. 14.46 on φ and v φ, and f 2 isolates the dependence of Eq. 14.47 on φ. Combining Eqs. 14.48 and 14.49 to eliminate z, and rearranging, yields

$$ 0 = \frac{{{{c}_{\sigma }}A}}{{{{v}_{\phi}}}}{{\left| {p_{\rm{e}}} \right|}^{{{{n}_{\sigma }} - 1}}}p_{\rm{e}}^{{}}{\hbox{d}}{{p}_{\rm{e}}} - \frac{{{{f}_1}}}{{{{f}_2}}}{\hbox{d}}{\phi}, $$
(14.50)

which must be satisfied by the φ-p e trajectory of any steady-state solution. Defining a function H such that

$$ H \equiv - \int {\frac{{{{f}_1}}}{{{{f}_2}}}{\hbox{d}}{\phi} }, $$
(14.51)

the integral of Eq. 14.50 yields a function

$$ U \equiv \frac{{{{c}_{\sigma }}A}}{{{{v}_{\phi}}}}\frac{{{{{\left| {p_{\rm{e}}} \right|}}^{{{{n}_{\sigma }} - 1}}}p_{\rm{e}}^{{^2}}}}{{{{n}_{\sigma }} + 1}} + H $$
(14.52)

whose φ-p e contours explicitly define the φ-p e trajectory for all steady-state solutions as a function v φ. Because U increases monotonically, and symmetrically, with positive or negative p e at constant φ, and H is independent of p e, the stationary points of U must occur at p e = 0 and correspond to extrema in H, i.e., the real roots of \( {{{\partial H}} \left/ {{\partial \phi = - {{{{{f}_1}}} \left/ {{{{f}_{{2}}}}} \right.} = 0}} \right.}. \) Moreover, as f 2 must be finite if the matrix is coherent, the roots of \( {{{\partial H}} \left/ {{\partial \phi = 0}} \right.} \) are identical to the roots of f 1 = 0. Therefore φ0 is always a stationary point, with the character of a focal point if \( {{{\partial {{f}_1}}} \left/ {{\partial \phi \, < \,0}} \right.} \) and that of a saddle point if \( {{{\partial {{f}_1}}} \left/ {{\partial \phi > \,0}} \right.} \). When φ0 is a focal point, the steady-state wave solutions correspond to periodic waves that oscillate between two values of porosity on either side of φ0, characterized by equal H, at which p e vanishes (Fig. 14.6b). The case of greater interest is a solitary wave (Fig. 14.6a), in which the porosity rises from φ0 to a maximum, at which Hmax) = H0), and then returns to φ0. This solution requires both the existence of a focal point at φ > φ0 and that φ0 is a saddle point. For the rheological constitutive relation employed here (Eq. 14.42), the first condition is always met when φ0 is a saddle point. Thus, the critical velocity for the existence of the solitary wave solution, i.e., the bifurcation at which φ0 switches from focal to saddle point, is

$$ v_{{\phi} }^{\rm{crit}} = - \frac{{{{k}_0}}}{\mu }\left( {1 - {{{\phi} }_0}} \right)\Delta \rho g\left( {\frac{{\left( {1 - {{{\phi} }_0}} \right)}}{{{{k}_0}}}{{{\left. {\frac{{\partial k}}{{\partial {\phi} }}} \right|}}_{{{\phi} = {{{\phi} }_0}}}} - 1} \right), $$
(14.53)

which is obtained by solving \( {{{\partial {{f}_1}}} \left/ {{\partial {\phi} = 0}} \right.} \) for v φ. Substituting the explicit function for permeability given by Eq. 14.17 into Eq. 14.53 yields

$$ v_{{\phi} }^{\rm{crit}} = - \frac{{{{k}_0}}}{{{{{\phi} }_0}\mu }}\left( {1 - {{{\phi} }_0}} \right)\Delta {\rho} g\left( {\left( {1 - {{{\phi} }_0}} \right){{n}_{\phi}} - {{{\phi} }_0}} \right) = {{v}_{{0}}}\left( {\left( {1 - {{{\phi} }_0}} \right){{n}_{\phi}} - {{{\phi} }_0}} \right). $$
(14.54)

Equation 14.54 implies that, in the small-porosity limit, the minimum speed at which steady solitary waves exist is n φ times the speed of the fluid through the unperturbed matrix.

The relation between solitary wave amplitude (maximum porosity) and v φ is obtained by solving

$$ H\left( {{{{\phi} }_{{\max }}}} \right) - H\left( {{{\phi }_0}} \right) = - \int\limits_{{{{\phi }_0}}}^{{{{\phi }_{{\max }}}}} {\frac{{{{f}_1}}}{{{{f}_2}}}{\hbox{d}}\phi } = 0. $$
(14.55)

The resulting expressions are cumbersome, but, in the small-porosity limit of Eqs. 14.17 and 14.42, the solution of Eq. 14.55 is

$$ {{v}_{\phi }} = - \frac{{{{c}_{\phi }}\phi_0^{{{{n}_{\phi }} - 1}}\Delta {\rho} {\text{g}}}}{\mu }\left( {{{n}_{\phi }} - 1} \right)\frac{{\phi_0^{{{{n}_{\phi }}}} + \phi_{{\max }}^{{{{n}_{\phi }}}}\left[ {{{n}_{\phi }}\ln \left( {\frac{{{{\phi }_{{\max }}}}}{{\phi_0}}} \right) - 1} \right]}}{{\phi_0^{{{{n}_{\phi }} - 1}}\left[ {{{n}_{\phi }}{{\phi }_{{\max }}} - \phi_0\left( {{{n}_{\phi }} - 1} \right)} \right] - \phi_{{\max }}^{{{{n}_{\phi }}}}}}. $$
(14.56)

From Eq. 14.56 it follows that n φ > 1 is a necessary condition for the existence of solitary waves. Equation 14.56 also has the surprising implication that amplitude is not a function of n σ, for large porosity the function f φ, in the exact form of Eq. 14.42, gives rise to a weak dependence of amplitude on n σ. For a solitary wave with specified phase velocity, the effective pressure is obtained as an explicit function of φ from the definite integral of Eq. 14.50, which can be rearranged to

$$ p_{\rm{e}} = \pm \root{{{n}_{\sigma}} + 1}\of{\left( {{n}_{\sigma} + 1} \right)\frac{{{{v}_{\phi}}}}{{{{c}_{\sigma}A}}\int\limits_{{{{{\phi} }_0}}}^{{\phi} } {\frac{{{{f}_1}}}{{{{f}_2}}}{\hbox{d}}{\phi} } }}, $$
(14.57)

where the signs have been dropped in view of the symmetry of the solution. And finally, substituting Eq. 14.57 into Eq. 14.47, inverting the result, and integrating yields the depth coordinate relative to the center of a wave as a function of φ

$$ z = \pm \sqrt[{{{n}_{\sigma }} + 1}]{{\frac{{{{v}_{\phi }}}}{{{{c}_{\sigma }}A{{{\left( {{{n}_{\sigma }} + 1} \right)}}^{{{{n}_{\sigma }}}}}}}}}{{\int\limits_{{{{\phi }_{{\max }}}}}^{\phi } {\frac{1}{{{{f}_2}}}\left( {\int\limits_{{{{\phi }_0}}}^{\phi } {\frac{{{{f}_1}}}{{{{f}_2}}}{\text{d}}\phi } } \right)} }^{{ - \frac{{{{n}_{\sigma }}}}{{{{n}_{\sigma }} + 1}}}}}{\text{d}}\phi . $$
(14.58)

To demonstrate that \( z \to \pm \infty \) as \( \phi \to {{\phi }_0} \), the inner integral and its factor in Eq. 14.58 are approximated by the first non-zero terms of Taylor series expansions about φ = φ0 to obtain

$$ z \approx \pm {{\left( {\frac{{{{v}_{\phi }}}}{{{{c}_{\sigma }}A{{{\left. {{{f}_2}} \right|}}_{{\phi = {{\phi }_0}}}}}}{{{\left( {\frac{{{{n}_{\sigma }} + 1}}{2}{{{\left. {\frac{{\partial {{f}_1}}}{{\partial \phi }}} \right|}}_{{\phi = {{\phi }_0}}}}} \right)}}^{{ - {{n}_{\sigma }}}}}} \right)}^{{\frac{1}{{{{n}_{\sigma }} + 1}}}}}\int\limits_{\Phi }^0 {{{\Phi }^{{ - \frac{{2{{n}_{\sigma }}}}{{{{n}_{\sigma }} + 1}}}}}{\hbox{d}}\Phi } $$
(14.59)

where Φ = φ − φ0. In the limit \( \Phi \to 0 \), the integral in Eq. 14.59 is finite only if n σ < 1, from which it is concluded that solitary waves have infinite wavelength in a linear or shear thinning viscous matrix, but may have finite wavelength in the peculiar case of a shear thickening viscous matrix. Rewriting the integral in Eq. 14.59 in terms of dlnφ, and differentiating yields

$$ \frac{{\partial z}}{{\partial \ln \Phi }} \approx {{\left( {{{\Phi }^{{1 - {{n}_{\sigma }}}}}\frac{{{{v}_{\phi }}}}{{{{c}_{\sigma }}A{{{\left. {{{f}_2}} \right|}}_{{\phi = {{\phi }_0}}}}}}{{{\left( {\frac{{{{n}_{\sigma }} + 1}}{2}{{{\left. {\frac{{\partial {{f}_1}}}{{\partial \phi }}} \right|}}_{{\phi = {{\phi }_0}}}}} \right)}}^{{ - {{n}_{\sigma }}}}}} \right)}^{{\frac{1}{{{{n}_{\sigma }} + 1}}}}} $$
(14.60)

the depth interval over which porosity decays from eφ0 to φ0 within a porosity wave. This interval is taken here as the characteristic length scale for variations in porosity, i.e., the viscous compaction length. The derivative on the right-hand side of Eq. 14.60,

$$ {{\left. {\frac{{\partial {{f}_1}}}{{\partial \phi }}} \right|}_{{\phi = {{\phi }_0}}}} = \Delta \rho g\left( {\frac{{\left( {1 - {{\phi }_0}} \right)}}{{{{k}_0}}}{{{\left. {\frac{{\partial k}}{{\partial \phi }}} \right|}}_{{\phi = {{\phi }_0}}}} - 1} \right) + \frac{{{{v}_{\phi }}\mu }}{{{{k}_0}\left( {1 - {{\phi }_0}} \right)}}, $$
(14.61)

is zero at \( {{v}_{\phi }} = v_{\phi }^{\rm{crit}} \), but decreases monotonically with v φ; thus dropping the first term in Eq. 14.61, substituting \( {{v}_{\phi }} = v_{\phi }^{\rm{crit}} \) and Φ = (e − 1) φ0 in Eq. 14.60, and expanding f 2 at φ0 as\( \left( {1 - {{\phi }_0}} \right){{\left. {{{f}_{\phi }}} \right|}_{{\phi = {{\phi }_0}}}} \)yields

$$ \delta = {{\left[ {{{{\left( {\left( {{\hbox{e}} - 1} \right){{\phi }_0}\Delta \rho {\hbox{g}}\left( {\left[ {1 - {{\phi }_0}} \right]{{{\left. {\frac{{\partial k}}{{\partial \phi }}} \right|}}_{{\phi = {{\phi }_0}}}} - {{k}_0}} \right)} \right)}}^{{1 - {{n}_{\sigma }}}}}{{{\left( {\frac{{2{{k}_0}}}{{{{n}_{\sigma }} + 1}}} \right)}}^{{{{n}_{\sigma }}}}}\frac{1}{{{{c}_{\sigma }}A\mu {{{\left. {{{f}_{\phi }}} \right|}}_{{\phi = {{\phi }_0}}}}}}} \right]}^{{\frac{1}{{{{n}_{\sigma }} + 1}}}}}, $$
(14.62)

a length scale that provides a lower bound on wavelength. For a linear-viscous matrix with shear viscosity η = 1/(3A), the constitutive relation given by Eq. 14.42, and the small-porosity limit, Eq. 14.62 simplifies to

$$ \delta = \sqrt{{\frac{4}{3}\frac{\eta }{{{\phi}_0}}\frac{{k_0}}{\mu}}} $$

which, accounting for differences in the formulation of the bulk viscosity of the matrix, is identical to the viscous compaction length of McKenzie (1984). For a non-linear viscous matrix, making use of constitutive relations given by Eqs. 14.17 and 14.42, in the small-porosity limit the compaction length is

$$ \delta = C\phi_0^{{\frac{{{{n}_{\phi }} - 1}}{{{{n}_{\sigma }} + 1}}}}\sqrt[{{{n}_{\sigma }} + 1}]{{{{{\left( {\frac{2}{{{{n}_{\sigma }} + 1}}} \right)}}^{{{{n}_{\sigma }}}}}\frac{{{{c}_{\phi }}}}{{{{c}_{\sigma }}A\mu {{{\left( {\Delta \rho {\text{g}}} \right)}}^{{{{n}_{\sigma }} - 1}}}}}}}, $$
(14.63)

where

$$ C = \root{{n}_{{\sigma}} + 1}\of{{{{{\left[ {{{n}_{\phi}}\left( {{\hbox{e}} - 1} \right)} \right]}}^{{1 - {{n}_{{\sigma}}}}}}}}. $$
(14.64)

The factor C represents two non-general assumptions of the analysis: that the phase velocity is n φ v 0; and that the porosity decay is from eφ0 to φ0. In the spirit of dimensional analysis, this factor (~2.27 for n σ = n φ = 3) is neglected in the text.

Rights and permissions

Reprints and permissions

Copyright information

© 2013 Springer Berlin Heidelberg

About this chapter

Cite this chapter

Connolly, J.A.D., Podladchikov, Y.Y. (2013). A Hydromechanical Model for Lower Crustal Fluid Flow. In: Metasomatism and the Chemical Transformation of Rock. Lecture Notes in Earth System Sciences. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-28394-9_14

Download citation

Publish with us

Policies and ethics