Alteration of the oceanic lithosphere and its implications for seafloor processes



Download 69.72 Kb.
Date31.03.2018
Size69.72 Kb.
#45177
Alteration of the oceanic lithosphere and its implications for seafloor processes

Wolfgang Bach1 and Gretchen L. Früh-Green2



Three-quarters of the global magmatism and 1/4th of the global heat loss are associated with tectono-magmatic and hydrothermal processes governing ocean lithosphere accretion and its aging from ridge to trench. Hydrothermal reactions between seawater and oceanic lithosphere under zeolite- to granulite-facies conditions are linked with magmatic and deformation processes but differ in nature relative to spreading rates.  Fast spreading ridges with frequent eruptions have telescoped metamorphic gradients and short-lived hydrothermal systems. The less magmatically-robust slow-spreading ridges are commonly cut by normal faults that expose ultramafic rocks on the seafloor and sustain long-lived hydrothermal systems with distinct vent fluid and fauna compositions.

Key words: ocean crust, water-rock interactions, geochemical fluxes, hydrothermal systems



  1. Geoscience Department, University of Bremen, 28359 Bremen, Germany, E-mail: wbach@uni-bemen.de

  2. Institute of Geochemistry and Petrology, ETH Zürich, CH-8092 Zürich, Switzerland, E-mail: frueh-green@erdw.ethz.ch

INTRODUCTION

On average approximately 3 km2 of new ocean crust is created annually along the >60,000 km long global network of mid-ocean ridges and back-arc spreading centers. This crust interacts with seawater throughout much of the lifetime of the seafloor (the average age of ocean crust is 60 million years). Fluid circulation and interaction with the crust is largely thermally driven and accounts for ¼ of the total global heat loss, or 11 TW (=Terawatts). Metamorphic conditions range from zeolite- to granulite-facies, with the degree of mass transfer being minimal to extensive. Hydrothermal circulation and interaction between seawater and the oceanic lithosphere are fundamental processes in controlling the composition of both seawater and the subducting oceanic plate. These processes play a critical role in regulating global heat and mass fluxes (German and Von Damm, 2003). They also determine geophysical and mineralogical properties of the oceanic lithosphere, with profound consequences for the mechanical behavior during aging and subduction. This is particularly relevant for volcanically starved mid-ocean ridge segments, where serpentinized mantle peridotite are denuded at the seafloor. The stability of serpentine and other hydrous phases determines the release of water in subduction zones. In submarine arc and back-arc hydrothermal settings located above subduction zones, silica- and volatile-rich magmas form as a consequence of the water flux into the mantle wedge. This, in turn, affects the composition of hydrothermal solutions, making arc-associated vent systems chemically and biologically distinct from their mid-ocean ridge counterparts. We discuss some of the principal processes controlling water-rock interaction in submarine hydrothermal systems, while highlighting some of the key differences in crustal architecture, magmatism and deformation between various geotectonic settings.



STRUCTURE OF THE OCEANIC LITHOSPHERE

Results of numerous oceanographic investigations and the ocean drilling programs (ODP and IODP) over the past three decades have changed the long-held view that the entire ocean crust is uniform in architecture and thickness, and have led to the recognition of fundamental differences in crustal accretion and alteration processes related to spreading rates (Figure 1). The principle of a uniform crust, characterized by a “layer-cake” structure of basalt, diabase (sheeted dike complex) and gabbro, is based on early geophysical data and comparisons with ophiolites, and is commonly referred to as the “Penrose ophiolite model” or “Penrose-type crust”, defined during a Penrose conference and published in Geotimes in 1972. This model is generally applicable for crust formed at fast spreading ridges, such as the East Pacific Rise, where melt supply is able to keep up with extension at the divergent plate boundary. However, it cannot be applied to much of the crust formed at oceanic ridges spreading at less than 40 mm yr–1 and which comprise nearly 50% of the global mid-ocean ridge system, extending from the Arctic Ocean, along the entire Mid-Atlantic Ridge (MAR) and into the SW Indian Ocean (SWIR). At these slow spreading ridges, melt supply does not always keep up with plate separation, which is accommodated by significant tectonic extension.

Recent investigations have shown that ultraslow spreading ridges, with spreading rates of <20 mm yr–1, have unique morphologies consisting of linked avolcanic and volcanic accretionary segments, with large variations in crustal thickness along ridge segments. Differences in melt volumes and focusing of melts are reflected by discontinuous volcanic centers separated by long stretches of avolcanic or melt-starved segments, in which the principal unit of accretion appears to be emplacement of mantle blocks onto the seafloor (Dick et al., 2003). Similarly, slow-spreading ridges, such as the Mid-Atlantic Ridge, are characterized by rugged topographies, a high degree of segmentation and considerable heterogeneities in crustal thickness, rock types, deformation and degree of alteration that reflect strong temporal and spatial variations in tectonic and magmatic processes (Figure 2). In these environments, serpentinized mantle peridotite and lower crustal plutonic rocks may represent 20-25% of the seafloor exposed through faulting (Cannat et al., 1997).

Using high-resolution multibeam bathymetry and side-scan sonar, recent studies have imaged broad, dome-like massifs with exceptional relief (>2.5 km) and with arched upper surfaces marked by distinctive spreading-direction-parallel corrugations and finer scale striations. These dome-shaped massifs are referred to as oceanic core complexes, in analogue to metamorphic core complexes on the continents, and form along long-lived detachment (normal) faults during phases of extensive tectonic extension, commonly associated with asymmetric spreading. Uplift and up to 40° rotation along these normal faults close to the ridge axis produce large individual domes or distinctive linear ridges with nearly symmetrical slopes. Detachment faults remove the basaltic upper crust and expose plutonic and upper mantle sequences from depths >6 km onto the seafloor in the footwalls (see Blackman et al., 2009, for a recent review) (Figure 2).

Core complexes have been found everywhere along segments at the SWIR and MAR (Cannat et al., 2006; Smith et al., 2008). For example, between the Fifteen-Twenty and Marathon fracture zones of the MAR, 45 potential core complexes, some less than 25 km apart, are scattered along two segments (13° and 15°N segments) and extinct core complexes are identified on the seafloor off-axis (Smith et al., 2006). The modern view of ocean crust at slow spreading ridges is that some portions of segments may consist of more uniform, Penrose-like crust, with characteristic abyssal hill morphologies, while other portions of the segments are highly heterogeneous with attenuated or missing volcanic crust and with gabbroic intrusions trapped in partially serpentinized lithospheric mantle (Figures 1 and 2). The depth and spatial extent of serpentinization of oceanic peridotites has direct implications for the interpretation of seismic velocities and may influence where geophysicists estimate the position of the Mohorovičić discontinuity (Moho). In these environments, the Moho may not represent the crust-mantle boundary but may be a serpentinization front.
Detachment faulting and channeled fluid flow

Oceanic core complexes and their exposed detachment faults represent tectonic windows that provide access to deep-seated rocks and allow studies of mantle flow, melt generation and migration, strain localization, and alteration. In these settings, variable amounts of gabbroic melts are trapped in the upper mantle, and detachment fault zones provide fluid pathways that may be important for cooling the lithosphere as well as for long-lived fluid circulation that can sustain hydrothermal systems (Figure 2c). In heterogeneous sections of the oceanic lithosphere, where gabbroic rocks are volumetrically variable and interdispersed with serpentinized peridotites, extensive talc-amphibole-chlorite metasomatism as well as heterogeneous, crystal-plastic to cataclastic deformation mark the zones of detachment faulting. Geochemical and microstructural studies indicate a complex mutual interaction between the gabbroic and ultramafic rocks at temperatures <500°C and suggest that these zones represent localized circulation of oxidizing, Si-Al-Ca-rich fluids and mass transfer during high strain deformation (Escartin et al., 2003; Boschi et al., 2006).

The presence of mechanically weak minerals, such as talc, serpentine and chlorite, may be critical to the development of detachment faults and may enhance unroofing during seafloor spreading (Escartin et al., 2003). Talc in particular may be influential in lubricating and softening mylonitic shear zones and can lead to strain localization and focused hydrothermal circulation for up to a few million years. Because of the long-lived nature of oceanic detachment faults, core complex development has been associated with the formation of the largest high-temperature, mineral-rich hydrothermal deposits found to date on the seafloor (McCaig et al., 2007).

Serpentinization processes

Mantle-dominated lithosphere is a highly reactive chemical and thermal system, in which interaction with seawater has significant geophysical, geochemical and biological importance for the global marine system and for subduction zone processes (Mével, 2003). Serpentinization involves hydration of olivine, the dominant mineral in the oceanic mantle, and typically incorporates >10 weight % H2O into the rocks. Hydration reactions are accompanied by a 20-40% increase in volume, a decrease in bulk density, and changes in mineralogy and rheology that directly affect the strength and physical properties of the mantle, the magnetic and gravity signatures, and the seismic velocities. Serpentinization processes also have major consequences for long-term, global geochemical fluxes by acting as a sink for H2O, Cl, B, U, S, C and a source of Ca, Ni and possibly Cr to hydrothermal fluids, and by producing extremely reduced fluids (Janecky and Seyfried, 1986; Allen and Seyfried, 2003; Früh-Green et al., 2004). In addition, seafloor weathering of serpentinized abyssal peridotites may result in Mg loss (Snow and Dick, 1995).

Since the earliest experimental studies of serpentines and studies of Alpine peridotites and ultramafic rocks in ophiolites, it has been known that serpentinization produces heat through exothermic reactions and conditions of low oxygen fugacity (fO2), resulting in the generation of H2-rich fluids and native metals. The formation of H2 during serpentinization is attributed to the production of magnetite during olivine hydration and is generally described by simplified model reactions with end-member phases. In reality, serpentinization involves solid solutions and metastable reactions governed by local variations in the activities of elements such as Si, Mg, Fe, Ca, and C (e.g., Janecky and Seyfried, 1986; Früh-Green et al., 2004). Hydrogen concentrations in the fluid may be limited by the formation of brucite (MgOH)2 and carbonates that can incorporate Fe2+ (Klein et al., 2009). Reduction accompanies the formation of FeNi alloys found in serpentinites, and a number of studies indicate that these alloys, magnetite, and/or Cr-bearing oxides may act as catalysts for the reduction of CO or CO2 to form abiotic methane and other hydrocarbons – a process referred to as Fischer-Tropsch-type reactions. The composition of any carbon-bearing fluid will thus be influenced by migration of the reduction front, and any CO2 or dissolved carbonate phases in the fluid may undergo reduction to hydrocarbons and graphite. The production of H2 and CH4 during serpentinization is also important in sustaining diverse microbial communities in subsurface and near-vent environments and may be significant for the existence of a deep biosphere (Früh-Green et al., 2004; Kelley et al., 2005).

The mineral assemblages and textures of oceanic serpentinites typically record progressive, static hydration reactions that take place under a wide range of temperatures, lithospheric depths, fluid compositions and redox conditions. The bulk volume of alteration is recorded by pseudomorphic mesh textures replacing olivine and bastite textures replacing pyroxene. Serpentinization is also accompanied by multiple generations of veins with a variety of vein-filling minerals, morphologies and textures that document local differences in formation mechanisms, stress regimes and fluid infiltration (Früh-Green et al., 2004; Andreani et al., 2007). Hydration expansion during serpentinization may also be a significant process in creating successive episodes of microfractures and in propagating cracks.

The products and sequence of serpentinization reactions depend on the depth, temperature and degree of seawater penetration into the lithosphere and will be influenced by bulk protolith compositions, the presence or absence of magmatic intrusions and/or trapped gabbroic melts, and structure (e.g., detachment faults, cataclastic fault zones). Thermodynamic constraints on phase equilibrium predict that high temperature serpentinization is sensitive to Si content and modal percentage of Opx (Allen and Seyfried, 2003; Klein et al., 2009). Depending on rock composition, serpentinization will commence once the shallow mantle sequences have cooled below 400-425°C, when olivine breakdown reactions to produce serpentine + magnetite are thermodynamically stable or below ~350° where serpentine + brucite are stable. Recent studies of serpentinites associated with detachment faults and core complexes indicate that serpentinization reactions in the footwall may initiate at depths of 4-6 km, which can be below the seismic Moho (Andreani et al., 2007). Early stages of high temperature seawater-mantle-gabbro interaction can also result in replacement of serpentine by talc-tremolite-chlorite assemblages to form fault rocks characteristic of detachment faults (Boschi et al., 2006). Progressive serpentinization reactions and veining continues until exposure on the seafloor. Pervasive serpentinization is likely to be restricted to the shallow lithosphere at depths <2 km, where brittle fractures and greater degrees of advective mass transport result in a transition to an open system of seawater infiltration and higher fluid/ratios. A progressive history of seawater-rock interaction in oceanic serpentinites and high integrated fluid/rock ratios are commonly recorded in shifts in oxygen, boron, strontium and neodymium isotope compositions (Boschi et al., 2006; Delacour et al., 2008).

MAGMATIC HYDROTHERMAL TRANSITION

The thermal and rheological structure of the oceanic crust is ultimately controlled by the balance between magmatic heat input and hydrothermal cooling (Phipps Morgan and Chen, 1993). During crustal accretion processes, this balance is determined by the complex interplay between magmatic, tectonic and hydrothermal processes. For example, although magmatic heat is the driving force for hydrothermal circulation of seawater in the oceanic crust, convective transport of heat influences the depths of crystallization and the cooling rates of magmas. Circulation is controlled by magmatic diking at fast spreading ridges, but may be tied more to extensional faults at slow spreading ridges (Wilcock and Delaney, 1996). Hydrothermal systems on the seafloor are small in size and short-lived at fast spreading ridges, while they are large and long-lived at slow spreading ridges. These differences are expressed in the style of seawater-rock interaction that can be observed in deep drill holes within the ocean crust and in ophiolites. Fast spread crust commonly shows progressive, static alteration controlled by fractures and microcracks, while in slow spread crust, more heterogeneous and extensive alteration is tied to ductile or brittle deformation in normal faults or to serpentinization processes in peridotite-dominated domains (Alt, 1995). In addition, the rheological behavior of the oceanic lithosphere, which likely plays a major role in the tectonic evolution of the ridge system, is determined by the hydration state of the solid crust and the presence of fluids.

The nature of the magmatic-hydrothermal interface in the oceanic crust is very poorly known, and we do not yet understand the mechanism by which heat is mined to drive mid-ocean ridge hydrothermal systems. Models of this interface fall into two major categories. One model assumes the presence of a conductive boundary layer between the hydrothermal cell and the heat source. Efficient heat transport in that conductive boundary layer is difficult to sustain, unless it is only a few meters thick (Cann et al., 1986). A second model, invoking a double-diffusive layer, assumes convective heat transport by brines that accumulate in the transition zone between hydrothermal and magmatic systems (Bischoff and Rosenbauer, 1989). These brines form by supercritical phase separation in the deep parts of hydrothermal systems at temperatures around 430-450˚C (Figure 3). It is commonly assumed that these temperatures are the maximum temperatures in the axial hydrothermal convection cells because of the rapid increase in the specific volume of the fluid above 400˚C. Large volumes of buoyant, vapor-rich fluids forming upon phase separation are expected to separate from the small volumes of brines. Yet other convection models do not invoke a brine layer and infer maximum fluid temperatures of 700˚C at the base of the circulation cell.  Such high fluid temperatures are consistent with temperatures obtained from plagioclase-amphibole thermometry for the uppermost gabbros from fast spread crust (Manning et al., 1996).

AXIAL HYDROTHERMAL SYSTEMS

From a rock perspective, hydrothermal alteration and mineralization patterns in ocean crust and ophiolite crust share both similarities and significant differences. In both settings, a conceptual model that divides a hydrothermal system into a recharge zone, a reaction (or root) zone, and a discharge (or upflow) zone can be applicable (Alt, 1995). Another commonality is the development of large sulfide deposits that form within the upper crust and at the seafloor due to heating of circulating seawater by crystallizing magmas and leaching of metals from the basement (e.g., Humphris and Cann, 2001; German and Von Damm, 2003). In ophiolites, however, the leaching zones near the sheeted dike-gabbro transition and in the deep upflow zones show abundant epidosites (base-metal depleted and Ca-metasomatized rocks), a rock type that has as yet not been recovered from in situ mid-ocean ridge sections (e.g., Alt, 1995). It also appears that greenschist-facies metamorphism is much more pervasive in the sheeted dikes and upper gabbros in ophiolites. The presence of epidosites and Sr and O isotopes provide evidence that fluid fluxes were greater in ophiolite sections, and it has been suggested that alteration patterns and styles in most ophiolite sections resemble forearc crust more than mid-ocean ridge crust (Gillis and Banerjee 2000). The upper greenschist-facies to lower amphibolite-facies mineral assemblages found in the lower sheeted dike complex in the deepest drill hole in ocean crust (ODP Hole 504B in the eastern equatorial Pacific) are consistent with a thermodynamic assessment of mineral-fluid equilibrium on the basis of vent fluid geochemistry (Bowers et al., 1985). The chemical composition of hydrothermal vent fluids implies temperatures of fluid-rock interaction around 420 to 440˚C (Seyfried et al., 1991). The conceptual model therefore is that black smoker vent fluids rise up from the root zones to the seafloor fairly rapidly and undergo cooling (adiabatic and conductively) by a few tens of degrees.



TAG

One of the largest and best-studied seafloor hydrothermal systems is the TAG hydrothermal mound at the MAR 26°N (Humphris at al., 1995). In 1994, the Ocean Drilling Program drilled several holes into the 200 m diameter and 50 m high sulfide mound, which hosts active black and white smoker complexes. Drilling revealed a complex internal stratigraphy, a dominance of breccias and a distinct mineralogical zonation, all of which are very similar to those observed in the feeder zones of the fossil hydrothermal system on the Galapagos Ridge and in massive sulfide deposits in ophiolites, such as the Troodos massif on Cyprus (Humphris at al., 1995; German and Von Damm, 2003). The upflow zone of hydrothermal fluids is zoned vertically and horizontally, with extensive massive brecciated pyrite ores at the top of the deposits, progressively grading into silica-anhydrite breccias that overlie an inner silica-pyrite-chlorite stockwork zone (Figure 4). The chlorite zone develops where upwelling hydrothermal solution mix with seawater that is entrained into the upflow zone. In the shallow and permeable parts of the system, mixing between hot hydrothermal fluids and seawater lead to precipitation of anhydrite and pyrite. Anhydrite plays a major role in the construction of the deposits and is precipitated through conductive heating of small amounts of entrained seawater, forming a structural framework during periods of high-temperature activity, followed by periods of inactivity in which anhydrite dissolves and causes the structure to collapse (Humphris et al., 1995). Fluids created by mixing of seawater and hydrothermal fluids have low pH and upon further migration through the sulfide mound, can remobilize elements like Zn, leading to zone refinement processes within the sulfide deposit and white-smoker type fluids venting on the seafloor. TAG is one of the largest sulfide deposits know in the oceans (4 Mio tons of sulfide with 30,000-60,000 tons of Cu) and is similar in size to ancient sulfide deposits in ophiolites.



Serpentinite-hosted systems

To date, five known active hydrothermal systems along the northern MAR are hosted in serpentinized peridotites with interdispersed gabbroic rocks and are located close to ridge discontinuities and segment ends: the Rainbow (36°N), Logatchev (15°N) and Ashadze (13°N), the Saldahna field (36°N), and the Lost City field (30°N). These hydrothermal systems vary in venting temperature, fluid compositions and type of hydrothermal deposits, but are characterized by high concentrations of H2, CH4, and other hydrocarbons, attributed to seawater interaction and serpentinization processes (Charlou et al., 2002; Kelley et al., 2005). The Rainbow, Logatchev and Ashadze fields are high-temperature (black smoker) systems with sulfide deposits deposited from low pH and Fe-rich fluids. Heat output of Rainbow has been estimated to be 1-5 GW (Thurnherr and Richard, 2001), making it the largest vent field yet discovered in terms of heat flux. High heat and fluid output, as well as the formation of sulfide deposits and high CO2 concentrations in the fluids, suggest that hydrothermal circulation in these high temperature systems is likely driven by yet-to-be-detected crystallizing magma bodies within or close to serpentinizing mantle peridotites.

The Lost City hydrothermal field is located on 1-1.5 my old serpentinized mantle rocks, 15 km west of the MAR axis and is distinctly different than all other known marine hydrothermal systems (Figure 5). At Lost City up to 60m high carbonate-brucite structures have formed where clear, high pH (from 9 to 11) fluids emanating from faults in serpentinite-dominated basement rocks exit on the seafloor. The fluids are up to 91°C and are depleted in metals and CO2, but have high concentrations of H2 and CH4 (and other low molecular weight hydrocarbons) that serve as important energy sources for anaerobic microorganisms within the porous chimney walls (Kelley et al., 2005; Proskurowski et al., 2008). Stable isotope and radiocarbon measurements on methane venting at Lost City have demonstrated that it is ultimately derived abiotically from mantle CO2 as opposed to seawater bicarbonate. This implies that seawater bicarbonate carried with recharge fluids is largely removed, presumably in carbonate minerals, before the abiotic reactions that form methane occur (Proskurowski et al., 2008).

Low to moderate temperature serpentinization reactions plays a key role in the production of high pH fluids and thus has important consequences for the sequestration of CO2 from seawater – an area of research that has stimulated considerable interest in industry and science communities. Investigations of modern-day serpentinization processes and the potential for CO2 sequestration through carbonate precipitation in peridotites have particularly gained importance as a possible technology for the reduction of anthropogenic CO2 input to the atmosphere (Kelemen and Matter, 2008).



OCEAN-CRUST EXCHANGE

From heat and geochemical budgets, one can derive that on the order of 1-4 TW of heat is extracted from the ridges to the oceans (Sleep, 1991; Elderfield and Schulz, 1996). This axial heat flux is only a small fraction of the total oceanic hydrothermal heat flux of 11 TW. The remaining heat is transported off-axis in ridge flank crust up to roughly 65 million years in age. The hydrothermal systems associated with these hydrologically active ridge flanks are low-temperature (<65°C), producing lower zeolite-facies alteration in the rocks.

The budget of Mg – one of the major cations in seawater – can only be explained when both types of hydrothermal circulation are considered (Mottl and Wheat, 1994). Mg is lost from seawater during interaction with ocean crust at all temperatures, but removal essentially becomes complete at temperatures above 100°C. On the order of several 1013 kg of seawater are fluxed through on-axis hydrothermal systems per year. The fluids become enriched in Ca, SiO2, alkalis, base metals, and sulfide but depleted in Mg, U, P, and sulfate. In general, the ocean crust is either a source or a sink for the elements dissolved in the oceans. Carbon dioxide, for example, is added to the oceans by magmatic degassing along mid-ocean ridges, but ridge flank alteration constitutes a sink of CO2 that is of about the same order of magnitude (1-2 x1012 moles/yr; Alt and Teagle, 1999). Likewise, alkali elements are leached from the rocks by seawater-derived fluids in high-temperature axial hydrothermal processes, while in the low-temperature ridge flank systems, they are lost from the circulating seawater to the ocean crust. The net effect here is that the crust is a prominent sink for alkali elements, and other elements like boron and uranium. Hydrous minerals (smectites, zeolites) and carbonates form in these ridge flanks systems and slowly seal the crust, which also becomes increasingly insulated from the oceans by accumulation of sediments. Both sediments and alteration minerals provide an important source of water and CO2 in subduction zones.

SEAFLOOR VENTS AND LIFE

The discovery of hydrothermal vents in the deep sea more than 30 years ago led to the now well-established recognition that geological processes along the 60,000 km long mid-ocean ridges support rich and diverse chemosynthetic ecosystems at the seafloor. In contrast to the ocean surface, water column, and marine sediments where essentially all life depends directly or indirectly on photosynthetic energy, vent communities are fuelled by chemical energy from the deep Earth (McCollom and Shock, 1997). There is an increasing awareness that magmatic processes and reactions between seawater and lithosphere support chemosynthetic ecosystems in many submarine geotectonic settings, such as mid-ocean ridges, intraplate volcanoes, forearcs, backarc basins, submarine arc volcanoes, continental margins and ridge flanks. What all these environments have in common are high energy potentials in form of redox disequilibria (Figure 6).

The basis of the food web in the deep oceans is chemosynthetic microorganisms that derive metabolic energy from the enzymatic catalysis of redox reactions. These organisms can thrive wherever reduced compounds, released by magma degassing or water–rock interactions, and oxidants mix within the ocean crust, at the seafloor, and in particle plumes originating from vent sites in the deep sea. Microorganisms can also mine energy from solids such as rocks, volcanic glass and sulfide minerals deposited at hydrothermal vents. Even at sites remote from spreading axes and thermal vents, chemosynthesis is now recognized to play a role in bare-rock systems of the deep ocean (Edwards et al., 2005).

In addition to stripping or leaching reduced components from magmas and rocks, geochemical energy is released where hydrogen is generated by hydrolysis during reaction of water with reduced Fe (or other metals) in basaltic glass or in Fe-rich minerals such as olivine. Dissolved H2 generated this way can be combined with CO2 or SO42- by methanogens or sulfate reducers, respectively. In serpentinizing environments, such as at Lost City, rare microbial communities of archaea and bacteria are supported by such hydrogen-rich vent fluids (Kelley et al. 2005; Brazelton et al., 2006). Many of these systems may host microbial communities and mineral-microbe interactions representing early metabolic pathways on Earth (Martin et al., 2008). There are many examples of how geochemical energy can be brought to the seafloor in fundamentally different ways as fluid flow regimes, magma supply, and rock composition vary within the oceanic lithosphere. Vent microorganisms can use this geochemical energy in a wide range of aerobic and anaerobic processes, many of which have not been fully explored to date.

ACKNOWLEDGMENTS

This contribution summarizes decades of research by the scientific community dedicated to studies of mid-ocean ridge processes, which to a large extent was conducted under the auspices of InterRidge and its individual country programs. The authors thank the numerous scientists who have contributed their knowledge to this subject, many of which could not be directly cited here. We also acknowledge the ocean drilling programs ODP and IODP and the participating scientists who have contributed greatly to a shift in paradigms and a better understanding of the formation, alteration, and aging of the oceanic lithosphere. We thank Mathilde Cannat and Håkon Austrheim for helpful comments. WB acknowledges support by the DFG-Research Center/Excellence in the cluster ‘The Ocean in the Earth System’, GFG acknowledges years of support by the SNF to conduct studies of serpentinization and the Lost City hydrothermal system.



FIGURE CAPTIONS

Figure 1: Two end-member models of crustal architecture at slow-spreading ridges. Although the seismic velocity structure is similar in both end-members, only one fits the common interpretation of such seismic profiles in terms of a layered magmatic crust. Note that in heterogenous crust, the Moho is a serpentinization front and not the crust-mantle boundary. Modified from Cannat (1993) in Mevel (2003).

Figure 2: Conceptual models of tectono-magmatic processes and their effects on alteration of the oceanic lithosphere associated with differences in spreading rates and based on geophysical and petrological constraints. (a) Robust magmatism at fast spreading ridges produces a relatively homogeneous, layered crust. Melts extracted from rising asthenosphere feed a narrow, sill-like melt lens 1-2 km beneath the ridge axis that grades downward into a partially crystallized mush zone, which is surrounded by a transitional zone of solidified by still hot gabbroic rock. Fluid circulation is limited to the crystallized upper crust and focused fluid circulation (dark red arrows) produce hydrothermal vent fields along the ridge axis. The relative volumes of melt and crystal mush vary along axis (b), especially near axis discontinuities (modified from Sinton and Detrick, 1992). In contrast, slow spreading ridges are characterized by variations in magmatic, tectonic and alteration processes along ridge segments. (c) Layered crustal sections form where magma supply is high, whereas detachment faults develop during extensive tectonic extension, where magma budgets are lower, and result in exposing mantle-dominated lithosphere to the seafloor (modified from Boschi et al., 2006). Much of the magma rising through this mantle section crystallizes at depth and forms variably sized gabbroic bodies. Detachment faults may intersect the crystallizing magma bodies and their feeder dikes and can channel high temperature fluids to feed large black smoker vent fields (dark red arrows). Serpentinization progresses as the mantle sections are uplifted and oceanic core complexes develop. Late normal faults and mass wasting can focus low temperature, alkaline hydrothermal fluids derived from serpentinization reactions (blue arrows), as seen at the Lost City vent field at the crest of the Atlantis Massif. (d) Along-strike variations in magma supply at slow spreading ridges are reflected by segment centers with relatively continuous, layered magmatic crust that becomes thinner and discontinuous toward magma-poor segment ends. Ultramafic rocks and serpentinization, with arrays of short strike slip and oblique faults, are common at segments ends and contribute to a thick lithosphere, as inferred from seismic data (modified from Cannat et al., 1995).

Figure 3: Phase relations in the H2O-NaCl system explaining the condensation of brines during supercritical phase separation (modified from Bischoff and Rosenbauer, 1984). The field for black smoker fluids encompassed many hundred observations (cf. German and Von Damm, 2003).

Figure 4: A schematic cross section through the TAG hydrothermal mound and the underlying stockwork zone. The basement underneath TAG reveals a vertical and concentric zoning typical of hydrothermal upflow zones, with a silicified and sulfidized central part and chloritized peripheral zones. Modified from Humphris et al. (1995).

Figure 5: Examples of (a) hydrothermal sulfide deposits and 360°C black smoker vents driven by magmatic activity (Logatchev hydrothermal field)(©marum) and (b) carbonate-brucite structures forming at moderate temperature (40-90°C) from high pH, metal-poor fluids emanating from serpentinized mantle peridotites at Lost City.

Figure 6. Cartoon depicting different geotectonic settings in which hydrothermal systems are developed and characterizing some of their principal differences. On axis systems comprise basalt- and peridotite-hosted vent fields, the latter of which are only a few km off-axis (figure is not to scale). Off-axis ridge flank systems are often related to seamounts that penetrate the impermeable sediment layer. Arc systems are influenced by CO2 and SO2 degassing of volatile-rich magma that reflect recycling of subducted components. The small inset cartoons illustrate some of the pronounced differences in substrate and vent fluid chemical compositions encountered in submarine vents around the globe.

REFERENCES

Allen DE, Seyfried WEJ (2003) Compositional controls on vent fluids from ultramafic-hosted hydrothermal systems at mid-ocean ridges: An experimental study at 400ºC, 500 bars. Geochimica Cosmochimica Acta 67(8):1531-1542

Alt JC (1995) Subseafloor processes in mid-ocean ridge hydrothermal systems. In: Humphris SE, Zierenberg RA, Mullineaux LS, Thomson RE eds) Seafloor hydrothermal systems, vol Geophysical Monograph. American Geophysical Union, Washington, DC, pp 85-114

Alt JC, Teagle DAH (1999) The uptake of carbon during alteration of ocean crust. Geochim. Cosmochim. Acta 63:1527-1535

Andreani M, Mevel C, Boullier AM, Escartin J (2007) Dynamic control on serpentine crystallization in veins: Constraints on hydration processes in oceanic peridotites. Geochemistry Geophysics Geosystems, 8:Q02012, doi:10.1029/2006GC001373.

Bischoff JL, Rosenbauer RJ (1984) The critical point and two-phase boundary of seawater, 200°-500°C. Earth Planet. Sci. Lett. 68:172-180

Bischoff JL, Rosenbauer RJ (1989) Salinity variations in submarine hydrothermal systems by layered double-diffusive convection. J. Geol. 97:613-623

Blackman DK, Canales P, Harding A (2009) Geophysical signatures of oceanic core complexes. Geophys. J. Int. 178:593-613

Boschi C, Früh-Green GL, Delacour A, Karson JA, Kelley DS (2006) Mass transfer and fluid flow during detachment faulting and development of an oceanic core complex, Atlantis Massif (MAR 30°N). Geochem. Geophys. Geosyst. 7:Q01004, doi:01010.01029/02005GC001074

Bowers TS, Von Damm KL, Edmond JM (1985) Chemical evolution of mid-ocean ridge hot springs. Geochim. Cosmochim. Acta 49:2239-2252

Brazelton WJ, Schrenk MO, Kelley DS, Baross JA (2006) Methane- and sulfur metabolizing microbial communities dominate the Lost City Hydrothermal Field ecosystem. Appl. Environ. Microbiol. 72:6257-6270

Cann JR, Strens MR, Rice A (1986) A simple magma-driven thermal balance model for the formation of volcanogenic massive sulphides. Earth Planet. Sci. Lett. 76:123-134

Cannat M (1993) Emplacement of mantle rocks in the seafloor at mid-ocean ridges. J. Geophys. Res. 98:4163-4172

Cannat, M, Mével, C, Maia, M, Deplus, C, Durand, C, Gente, P, Agrinier, P, Belarouchi, A, Dubuisson, G, Humler, E, Reynolds, J (1995) Thin crust, ultramafic exposures, and rugged faulting patterns at the Mid-Atlantic Ridge (22°-24°N). Geology 23:49-52

Cannat M, Lagabrielle Y, de Coutures N, Bougault H, Casey J, Dmitriev L, Fouquet Y (1997) Ultramafic and gabbroic exposures at the Mid-Atlantic Ridge: geological mapping in the 15°N region. Tectonophysics 279:193-213

Cannat M, Sauter D, Mendel V, Ruellan E, Okino K, Escartin J, Combier V, Baala M (2006) Modes of seafloor generation at a melt-poor ultraslow-spreading ridge. Geology 34:605-608

Delacour A, Früh-Green GL, Frank M, Gutjahr M, Kelley DS (2008) Sr- and Nd-isotope geochemistry of the Atlantis Massif (30°N, MAR): Implications for fluid fluxes and lithospheric heterogeneity. Chemical Geology 254:19-35

Dick HJB, Lin J, Schouten H (2003) An ultraslow-spreading class of ocean ridge. Nature 426:405-412

Edwards KJ, Bach W, McCollom TM (2005) Geomicrobiology in Oceanography: Mineral-Microbe Interactions in the Deep-Sea. Trends in Microbiology 13(9):449-456

Elderfield H, Schultz A (1996) Mid-ocean ridge hydrothermal fluxes and the chemical composition of the ocean. Annu. Rev. Earth Planet. Sci. 24:191-224

Escartin J, Mevel C, MacLeod CJ, McCraig AM (2003) Constraints on deformation conditions and the origin of oceanic detachments: The Mid-Atlantic Ridge core complex at 15º45'N. Geochem. Geophys. Geosys. 4(8):10.1029/2002GC000472

Früh-Green GL, Connolly JAD, Plas A, Kelley DS, Grobety B (2004) Serpentinization of oceanic peridotites: implications for geochemical cycles and biological activity. In: Wilcock WSD, DeLong EF, Kelley DS, Baross JA, Cary SC (eds) The subsurface biosphere at mid-ocean ridges, vol Geophysical Monograph. American Geophyscial Union, Washington, DC, pp 119-136

German CR, Von Damm, KL (2003) Hydrothermal Processes. In: Holland HD, Turekian KK (eds) Treatise on Geochemistry, Elsevier, Amsterdam

Gillis KM, Banerjee NR (2000) Hydrothermal alteration patterns in supra-subduction zone ophiolites. In: Dilek Y, Moores EM, Elthon D, Nicolas A (eds) Ophiolites and Oceanic Crust: New insights from field studies and the Ocean Drilling Program, vol 349. Geological Society of America, Special Paper, Boulder, CO, pp 283-297

Humphris SE, Cann. JR (2000) Constraints on the energy and chemical balances of the modern TAG and ancient Cyprus seafloor sulfide deposits. J. Geophys. Res. 105(12):28,477-428,488

Humphris SE, Herzig PM, Miller DJ, Alt JC, others a (1995) The internal structure of an active sea-floor massive sulphide deposit. Nature 377:713-716

Janecky DR, Seyfried WE, Jr. (1986) Hydrothermal serpentinization of peridotite within the oceanic crust: Experimental investigations of mineralogy and major element chemistry. Geochim. Cosmochim. Acta 50:1357-1378

Kelley DS, Karson JA, Fruh-Green GL, Yoerger DR, Shank TM, Butterfield DA, Hayes JM, Schrenk MO, Olson EJ, Proskurowski G, Jakuba M, Bradley A, Larson B, Ludwig K, Glickson D, Buckman K, Bradley AS, Brazelton WJ, Roe K, Elend ML, Delacour A, Bernasconi SM, Lilley MD, Baross JA, Summons RE, Sylva SP (2005) A serpentinite-hosted ecosystem: The Lost City Hydrothermal Field. Science 307:1428-1434

Klein F, Bach W, Jöns N, McCollom TM, Moskovitz B, Berquo T (2009) Iron partitioning and hydrogen generation during serpentinization of abyssal peridotites from 15°N on the Mid-Atlantic Ridge. Geochim Cosmochim Acta 73: 6868-6893

Manning CE, Weston PE, Mahon KI (1996) Rapid high-temperature metamorphism of East Pacific Rise gabbros from Hess Deep. Earth Planet. Sci. Lett. 144:123-132

McCaig AM, Cliff RA, Escartin J, Fallick AE, MacLeod CJ (2007) Oceanic detachment faults focus very large volumes of black smoker fluids. Geology 35:935-938

McCollom TM, Shock EL (1997) Geochemical constraints on chemolithoautotrophic metabolism by microorganisms in seafloor hydrothermal systems. J. Geophys. Res. 61:4375-4391

Mottl MJ, Wheat CG (1994) Hydrothermal circulation through mid-ocean ridge flanks: Fluxes of heat and magnesium. Geochim. Cosmochim. Acta 58:2225-2237

Phipps Morgan J, Chen YJ (1993) The genesis of oceanic crust: Magma injection, hydrothermal circulation, and crustal flow. J. Geophys. Res. 98:6283-6298

Proskurowski G, Lilley MD, Seewald JS, Früh-Green GL, Olson EJ, Lupton JE, Sylva SP, Kelley DS (2008) Abiogenic Hydrocarbon Production at Lost City Hydrothermal Field. Science 319:604-607

Seyfried WE, Ding K, Berndt ME (1991) Phase equilibria constraints on the chemistry of hot spring fluids at mid-ocean ridges. Geochim. Cosmochim. Acta 55:3559-3580

Sinton JM, Detrick RS (1992) Mid-ocean ridge magma chambers. J. Geophys. Res. 97:197-216

Sleep NH (1991) Hydrothermal circulation, anhydrite precipation, and thermal structure at ridge axes. J. Geophys. Res. 96:2375-2387

Smith DK, Cann JR, Escartin J (2006) Widespread active detachment faulting and core complex formation near 13°N on the Mid-Atlantic Ridge. Nature 442:440-443

Snow JE, Dick HJB (1995) Pervasive magnesium loss by marine weathering of peridotite. Geochim. Cosmochim. Acta 59:4219-4235



Thurnherr AM, Richards KR (2001) Hydrographic and high-temperature heat flux of the Rainbow hydrothermal site (36°14′N, Mid-Atlantic Ridge). J. Geophys. Res. 106:9411-9426


Download 69.72 Kb.

Share with your friends:




The database is protected by copyright ©ininet.org 2024
send message

    Main page