U.S. Dept. of Commerce / NOAA / OAR / PMEL / Publications


Hydrothermal Plumes Over Spreading-Center Axes: Global Distributions and Geological Inferences

Edward T. Baker

NOAA Pacific Marine Environmental Laboratory, Seattle, Washington, USA

Christopher R. German

Institute of Oceanographic Sciences, Wormley, Surrey, UK

Henry Elderfield

Department of Earth Sciences, Cambridge University, Cambridge, UK

Seafloor Hydrothermal Systems: Physical, Chemical, Biological, and Geological Interactions, Geophysical Monograph 91, AGU, 47-71, 1995.
Copyright ©1995 by the American Geophysical Union. Further electronic distribution is not allowed.

ABSTRACT

Hydrothermal plumes, formed by the turbulent mixing of hot vent fluids and ambient seawater, are potent tools for locating, characterizing, and quantifying seafloor hydrothermal discharge. Hydrographic, optical, and chemical tracers have all been used successfully to identify plumes. Observational techniques have progressed from discrete samples collected on vertical casts to continuous, in situ physical and chemical analyses that produce two- and three-dimensional plume maps. We present here a synthesis of available data from spreading centers throughout the world ocean wherever plumes have been mapped on a vent-field, segment, or multisegment scale. About 3% of the global divergent plate margin has been scrutinized for hydrothermal plumes; close to twice that distance has been sampled at least cursorily. Along medium- to superfast-spreading ridges in the eastern Pacific, where the most detailed work has been done, venting is common and plumes overlie 20-60% of the ridge crest length. Plumes are found also wherever careful surveys have been conducted in western Pacific marginal basins. Slow-spreading ridges in the North Atlantic, sampled over greater length scales but in less detail than in the Pacific, appear to have a comparatively low incidence of venting. Little is known of plume distributions over ridges in other oceans. These studies confirm that hydrothermal venting is present across the entire range of spreading rates and that continuous, underway plume surveys are the most efficient means available for locating seafloor discharge sites. Moreover, plume surveys are the only practical approach to mapping hydrothermal discharge patterns over multisegment spatial scales, and to integrating hydrothermal fluxes on a vent-field or larger scale. The plume surveys conducted to date indicate that the incidence of hydrothermal plumes over the ridge axis increases directly with increasing spreading rate. This observation supports models of crustal evolution that predict a direct relationship between the axial hydrothermal heat flux and spreading rate. This conclusion must be tempered, however, by the recognition that most of the global spreading center system remains unexplored for hydrothermal activity.

INTRODUCTION

Seafloor hydrothermal circulation is the principal agent of energy and mass exchange between the ocean and the earth's crust. Discharging fluids cool hot rock, construct mineral deposits, nurture biological communities, alter deep-sea mixing and circulation patterns, and profoundly influence ocean chemistry and biology. Although the active discharge orifices themselves cover only a minuscule percentage of the ridge-axis seafloor, the investigation and quantification of their effects is enhanced as a consequence of the mixing process that forms hydrothermal plumes. Hydrothermal fluids discharged from vents are rapidly diluted with ambient seawater by factors of 104-105 [Lupton et al., 1985]. During dilution, the mixture rises tens to hundreds of meters to a level of neutral buoyancy, eventually spreading laterally as a distinct hydrographic and chemical layer with a spatial scale of tens to thousands of kilometers [e.g., Lupton and Craig, 1981; Baker and Massoth, 1987; Speer and Rona, 1989].

Early investigations focused on plumes simply as indicators of nearby discharge sources. More comprehensive studies, however, quickly demonstrated that plumes integrate hydrothermal heat and mass flux [Baker and Massoth, 1987; Rosenberg et al., 1988; Rudnicki and Elderfield, 1992], provide natural laboratories for measuring the chemical reactions that control the ultimate dispersal of various hydrothermal species [Klinkhammer et al., 1983; Mottl and McConachy, 1990; Feely et al., 1991; German et al., 1991; Metz and Trefry, 1993; Rudnicki and Elderfield, 1993; Kadko, 1994], disperse vent-specific biological populations [Mullineaux et al., 1995], track deep-ocean circulation patterns [Lupton and Craig, 1981; Reid, 1982], and, in general, provide opportunities for identifying and quantifying the effects of hydrothermal activity. In this paper we focus narrowly on the use of plumes as a tool for hydrothermal exploration and geological interpretation. We begin by outlining commonly used techniques for producing detailed maps of the hydrographic and chemical anomalies produced by plumes along oceanic spreading centers. We then review those ridge-crest areas in the Pacific (including the western Pacific marginal basins), Atlantic, and Indian Oceans where sufficient data exist to map, or at least detect, plumes at vent-field, segment, or multisegment scales. Many of these data are new or formerly available only in fragmented form. Finally, we offer some geological interpretations of the data, using as an example the emerging coherence between hydrothermal discharge patterns and ridge-crest spreading rate.

OBSERVATIONAL TECHNIQUES

Formation of Plumes

Hydrothermal plumes form above sites of venting because of the buoyancy of the hot hydrothermal fluids that rise, entraining ambient sea water with a consequent continuous increase in plume volume, until neutral buoyancy is achieved and the plume disperses laterally. Models describing the physical and chemical properties of hydrothermal plumes have been presented by several workers [e.g., Middleton and Thomson, 1986; Little et al., 1987; Speer and Rona, 1989; McDougall, 1990; Rudnicki and Elderfield, 1992; Lavelle, 1994], all based upon the principles for turbulent entrainment in plumes described by Morton et al. [1956] and Turner [1973]. Using these models it is possible to calculate heat fluxes from plume and local hydrographic data, and to predict the ratio of entrained seawater to hydrothermal fluid as a function of plume height. During plume rise, ambient seawater is entrained from a range of depths such that the concentration of a property in the neutrally buoyant plume is a function of both the concentration and flux of that property in the vent fluid, and of the background profile of the property integrated over the whole height of rise of the buoyant plume. These complexities are discussed more fully in the following sections, and especially in accompanying papers by McDuff [this volume], Lupton [this volume], and Helfrich and Speer [this volume].

Hydrographic and Optical Tracers

Quantification of hydrothermal temperature anomalies is complicated by the variable background profiles of temperature and salinity found in the world ocean. In the deep Pacific, the ambient salinity gradient increases with depth and so the neutrally buoyant plume is warmer and saltier than its surroundings. An estimate of the hydrothermal temperature anomaly () is calculated as

                             (1)

where k and b are the slope and intercept, respectively, of the generally linear relationship between potential temperature () and potential density () in hydrothermally unaffected water immediately above the neutrally buoyant plume. To properly compare this value in the neutrally buoyant plume with the hydrothermal heat input at the seafloor source, however, the plume values must be corrected for the effect of local hydrography and vent fluid salinity.

In the Atlantic and wherever else the salinity gradient is negative above the ridge crest, the neutrally buoyant plume is cooler and fresher than the surrounding seawater. Equation (1) gives a negative value in this case, and hydrographic mapping of the plume is problematic at best. Evaluation of temperature anomalies must be considered within the framework of a plume model [e.g., Speer and Rona, 1989; McDougall, 1990; Rudnicki and Elderfield, 1992]. Thus most plume mapping in the Atlantic has depended on optical or chemical indices of hydrothermal discharge.

Both light attenuation and light scattering have been widely used to define hydrothermal plumes. Unlike conservative hydrographic tracers, optical tracers are nonconservative and depend on an ever-changing balance between particle production by precipitation and biological growth, and particle loss by deposition, dissolution, and ingestion. Hydrothermal fluids with a very low level of precipitable species, such as the vapor phase of a phase-separated fluid, or a fluid discharging through a thick layer of sediments, may yield a plume with a tenuous optical signal, but optically invisible plumes are rare. The only documented case we are aware of is a CH4 plume with no nephelometer or total dissolved Mn signal near 15°N on the Mid-Atlantic Ridge (MAR), a plume hypothesized to arise from fluid circulation in ultrabasic (low silica) rocks rather than basalt [Charlou et al., 1991a]. Moreover, optical tracers are often more sensitive and simpler to interpret than hydrographic tracers, since their background profiles are roughly vertically uniform. Their nonconservative nature can add to their usefulness, since an optical signal decreases more sharply than temperature with distance from its source.

Chemical Tracers

Hydrothermal fluids are enriched by up to a factor of 107 in several key tracers (e.g., Mn, Fe, CH4, H2, 3He, and many other trace species) relative to typical oceanic deep waters, so chemical anomalies associated with hydrothermal plumes can often be detected at significant distances away from hydrothermal vent sites. (The chemistry of hydrothermal fluids and plumes is more fully discussed by Lilley et al. [this volume], Lupton [this volume], Kadko et al. [this volume], and Von Damm [this volume].) Dissolved manganese (Mn) and methane (CH4) are commonly used to define hydrothermal plumes for three reasons. First, both are enriched ~106-fold in high-temperature vent fluids relative to ambient seawater [e.g., Welhan and Craig, 1983; Von Damm, 1990], so their concentrations in neutrally buoyant hydrothermal plumes directly above hydrothermal vent sites are ~102-fold higher than typical oceanic deep water. Second, both can be analyzed precisely at sea. Third, the contrasting quasiconservative behavior of Mn and distinctly nonconservative behavior of CH4 provide dynamical information about plumes. Because neutrally buoyant plumes overlie a much greater area of the midocean ridge crest than is occupied by active hydrothermal chimneys and mounds, these water column enrichments present geochemists a magnifying glass with which to prospect for new hydrothermal vent sites.

Fig. 1. Schematic comparison of three methods of shipboard plume mapping and sampling.

Sampling Techniques

Hydrographic and optical measurements are commonly combined in the form of a conductivity/temperature/depth (CTD) system interfaced with a transmissometer or nephelometer (or both). A basic sampling strategy is simply vertical casts widely spaced along the ridge axis (Figure 1). Because hydrothermal plumes are dynamic features with steep horizontal and vertical gradients, however, it is useful to adopt a sampling scheme that will reduce the problem of temporal and spatial aliasing as much as possible. Data return should be maximized from the plume horizon and minimized from waters above the plume where little useful information is present. An efficient strategy in mapping and sampling neutrally buoyant plumes is thus to conduct lengthy "tow-yos," continual raisings and lowerings of the instrumentation through the plume while slowly steaming (Figure 1). Experience shows that a typical CTD package can be towed with standard electromechanical cable at a rate of about 3-4 km hr-1, resulting in a complete up-down cycle through a 400-m plume layer about every 1 km along track using a winch speed of ~1 m s-1.

The majority of geochemical studies associated with hydrothermal plume surveys up to about 1990, and which continue in widespread use today, required conventional collection of seawater samples using CTD-rosette systems followed by ship-board or shore-based analyses for geochemical tracers. To accommodate the enlarged data sets required to address more complex geochemical problems, much investment has been made recently in novel instrumentation which can achieve greater sampling coverage or even produce real-time in situ geochemical analyses. This instrumentation will generate data in a form comparable to that readily obtainable from hydrographic and optical sensors, thus facilitating improved understanding of the biogeochemical processes active in plumes.

The first advance in geochemical studies of hydrothermal plumes came from the IFREMER laboratory's development of a continuous underway towed water sampler--the "Palanquée Dynamique" or "Dynamic Hydrocast" [Bougault et al., 1990]. This system comprises four sets of water sampling rosettes, each with a CTD sensor package and data logger, that are clamped to a cable and towed at different, known heights above the seafloor (Figure 1). Each set of water sampling rosettes comprises ten water-sampling bottles which fill progressively, and sequentially, as the rosette package is towed through the water column. The gearing on the rosettes can be adjusted so that each bottle fills progressively over 1-2 km, thus from a single deployment a series of ten integrated water samples are collected from each of four depths above the seabed over a total along-axis length of 10-20 km.

Although the Dynamic Hydrocast system enables systematic and continuous water-sampling, only a limited number of samples can be collected and these must be analyzed onboard or ashore. An alternative approach was taken by Johnson and co-workers at the Moss Landing Marine Laboratories. They developed a novel instrument package, the Submersible Chemical Analyzer (SCANNER), that could carry out analyses of nutrients, sulfide, Mn and Fe in situ [Johnson et al., 1986a,b; Coale et al., 1991; Chin et al., 1994]. The system is a deep-towed spectrophotometer that measures dissolved Mn and total dissolved Fe (Fe(II) and Fe(III)) every 5 seconds in seawater that has previously been passed through a 10 m mesh prefilter. The system connects routinely to a conventional CTD package augmented with a nephelometer and/or transmissometer. Coale et al. [1991] and Chin et al. [1994] used the SCANNER to map dissolved Fe and Mn anomalies in the range 20-50 nmol l-1 Fe and 50-200 nmol l-1 Mn in plumes over the Cleft segment of the Juan de Fuca Ridge.

More recently, SUAVE (Submersible System Used to Assess Vented Emissions), an enhanced version of the SCANNER system, has been developed by NOAA's Pacific Marine Environmental Laboratory. The SUAVE system comprises a total of six on-line colorimetric chemical detectors for Mn, Fe(II), Fe(II+III), Si, H2S, and one of PO4 or Cl, currently under development for plume and diffuse-flow studies, respectively [Massoth et al., 1991] (G. J. Massoth, personal communication, 1994). The system can currently be deployed for at least ~80 hours. Over 800 hours of successful deployment have been achieved to date upon a variety of platforms, including the manned submersible Alvin, the unmanned remotely operated vehicle ROPOS, and conventional towed CTD packages.

Sensitivity of both the SCANNER and SUAVE systems are limited by the colorimetric techniques available for dissolved Mn and Fe analyses, originally ~20 nmol l-1 and 25 nmol l-1 respectively [Chin et al., 1994]. New techniques have improved resolution to ~5-10 nmol l-1 (G. J. Massoth, personal communication, 1994). This sensitivity is sufficient for most, but not all, known plume anomalies. Dissolved Mn anomalies of only 15-20 nmol l-1 have been measured at the Snakepit and Broken Spur vent fields at 23°N and 29°N MAR [Elderfield et al., 1993; James et al., in press a], near the resolution level of the SCANNER or SUAVE systems.

To overcome this potential problem, Klinkhammer [1994] has developed an even more sensitive dissolved Mn analyzer, ZAPS (Zero Angle Photon Spectrophotometer), that can detect dissolved Mn to ambient seawater concentrations (1 nmol l-1), and can thus be employed to prospect for and investigate neutrally buoyant hydrothermal plumes anywhere on the world's spreading center system. ZAPS is a fiber-optic spectrometer which combines solid-state chemistry with photomultiplier tube detection to make flow-through chemical measurements in situ. ZAPS is also capable of being deployed as a light scattering device or ultraviolet fluorimeter [Klinkhammer et al., 1994].

GLOBAL DISTRIBUTIONS

Precursors of Modern Plume Surveys

Even before the discovery of submarine hydrothermal vents in 1977 [Corliss et al., 1979] and 1979 [RISE Project Group, 1980], a few investigators had recognized large-scale physical and chemical anomalies that were rightly attributed to hydrothermal discharge. Over 30 years ago, Knauss [1962] suggested that abnormally warm water between 3000 and 3500 m over the East Pacific Rise (EPR) near 20°S was probably related to a recently mapped zone of high seafloor heat flow in the same area [Von Herzen, 1959]. Ten years later, Warren [1973] noticed the same feature in data from the SCORPIO expedition at 28°S and 43°S, although he was reluctant to attribute the anomaly to geothermal effects since purely oceanographic effects might also lead to the same temperature distribution. Lonsdale [1976] made perhaps the earliest attempt to quantify the temperature anomaly of a hydrothermal plume when he plotted temperature-salinity curves of profiles west, east, and directly over the EPR at 8°S. The axial stations revealed near-bottom temperature increases of several hundredths of a degree Celsius relative to off-axis water of the same salinity (Figure 2).

Fig. 2. Plot of potential temperature vs. salinity for discrete samples near the East Pacific Rise at 8°S showing hydrothermal heating at the rise crest (reprinted from Lonsdale [1976]).

Measurements of the helium isotopes 3He and 4He in the Kermadec Trench [Clarke et al., 1969] and across the EPR [Craig et al., 1975] provided early evidence that hydrothermal venting could produce widely dispersed chemical anomalies. Shortly thereafter, sampling near the Galapagos Ridge by Klinkhammer et al. [1977] and Bolger et al. [1978] showed that hydrothermal plumes also contain significant anomalies of dissolved and particulate Mn. Since then, plume surveys have contributed to the discovery and characterization of hydrothermal discharge sites on ridges and hotspots throughout the oceans and marginal seas.

Eastern Pacific Spreading Centers

Hydrothermal plumes have been mapped in more detail over eastern Pacific ridge crests than anywhere else in the ocean. Virtually continuous surveys of plume hydrographic, optical, and/or chemical anomalies have been conducted along much of the Explorer-Juan de Fuca-Gorda Ridge complex in the northeast Pacific, in the Gulf of California, and along the EPR from about 9° to 13°N and 13° to 22°S. Full-rate spreading speeds at these sites range from a medium rate of 60 mm yr-1 in the northeast Pacific and the Gulf of California, to intermediate rates of 110-120 mm yr-1 near 10°N, to superfast rates of >150 mm yr-1 near 20°S [DeMets et al., 1990]. The comprehensive nature of plume maps in the eastern Pacific results in large part from the narrowness of the axial ridge crest at medium and greater spreading rates, which allows a single along-axis transect to detect and locate most discharge sources.

Fig. 3. Location map for northeast Pacific spreading centers.

The Explorer-Juan de Fuca-Gorda Ridge complex is an isolated set of spreading ridges that extend for almost 1000 km along the west coast of Canada and the United States (Figure 3). The Juan de Fuca Ridge (JDFR) has been surveyed extensively, while Explorer and Gorda Ridges have benefitted only from site-specific work. McConachy and Scott [1987] used real-time nephelometer signals to map the distribution of plumes over a 40 km2 area of Southern Explorer Ridge. Specific plume maxima were linked to two known vents sites near 49°46'N, 130°16'W, plus undiscovered vent fields farther north. On the Gorda Ridge, plumes from high-temperature venting appear restricted to the northern third of the axial valley [Baker et al., 1987a; Collier and Baker, 1990]. Discharge from vents within the thickly sedimented Escanaba Trough in the southern third of the Gorda Ridge produces plumes difficult to detect chemically, since conductive cooling and precipitation removes most of the dissolved metals during passage through the sediments [Campbell et al., 1994; German et al., in press].

Plumes emanating from the JDFR have been mapped more intensively and extensively than those over any other group of ridge-crest segments in the ocean. The first indication of JDFR hydrothermal plumes came in 1980 from observations of dissolved Mn [Jones et al., 1981] and 3He [Lupton, 1990] at five widely spaced stations stretching from 44°31'N to 47°47'N. In 1982, Crane et al. [1985] conducted the first large scale, continuous survey of hydrothermal plumes anywhere on the global midocean ridge (MOR). They attempted to measure hydrothermally induced temperature anomalies in plumes by hanging a 50-m-long thermistor chain from a side-scan sonar vehicle while towing it along the entire JDFR. Their measurements were successful in identifying a few major thermal anomalies but did not fully reveal the actual plume distribution because of calibration problems and the very slow (~2 min) response time of the thermistors. Nevertheless, the methods and strategy of Crane et al. [1985] were precursors for many studies that later used two-dimensional transects of temperature and chemical anomalies to approximately locate vent sources, estimate hydrothermal heat and mass fluxes, and relate spatial and temporal variability of venting to the magmatic budget of the ridge.

Plume mapping on the JDFR has focused on three areas: the Endeavour segment, Axial Volcano, and the Cleft segment (Figure 3). Along the Endeavour segment, Baker and Massoth [1987] and Thomson et al. [1992] used transects of temperature and light attenuation anomalies to characterize the plume and discover vent locations. They combined these anomalies with concomitantly collected records of plume advection from moored current meters to estimate mean heat fluxes of 1700 ± 1000 MW [Baker and Massoth, 1987] and 800 ± 490 MW [Thomson et al., 1992]. Thomson et al. [1992] also reported an "instantaneous" heat flux of 12,000 ± 6000 MW based on temperature and plume velocity measurements directly over the central vent site. Rosenberg et al. [1988] combined temperature anomaly profiles with discrete sampling of 222Rn and 3He to make a geochemical estimate of heat flux based on the premise that the loss of 222Rn from the plume by decay is exactly balanced by the constant addition of hydrothermal 222Rn. Their best estimate of heat flux was 3000 MW. Although the 3He plume from the summit of Axial Volcano is traceable for tens of kilometers [Lupton, 1990], temperature and light attenuation anomalies are detectable only within the summit caldera. Baker et al. [1990] used temperature anomaly and plume advection measurements to estimate a heat flux 800 MW from the caldera.

Fig. 4. CTD tow-yo transects of hydrothermal temperature anomaly (°C) along the Cleft segment axial valley, 1986-1991. Triangles show location of known venting sites. The lower half of event plume EP86 ("megaplume") is apparent in the 1986 transect. The anomaly at 1800 m in 1988 is a hydrographic rather than hydrothermal feature [Baker, 1994].

The plume distribution over the entire Cleft segment has been mapped one or more times every year since 1986 [Baker and Massoth, 1986; Baker and Massoth, 1987; Baker, 1994], making it a unique source of information about the temporal variability of hydrothermal venting. Discharge from the Cleft segment is concentrated in two areas centered at 44°38'-44°43'N and 44°54'-45°N (Figure 4). The first plume survey to estimate hydrothermal fluxes was conducted at the southern site in 1985 [Baker and Massoth, 1986; 1987], finding a heat flux of 580 ± 350 MW. A six-year average of plume extent and temperature anomaly at the northern site, combined with long-term current meter measurements of advection at plume depths [Cannon et al., 1991], yielded a flux estimate of 660 ± 280 MW [Baker, 1994]. Continuous monitoring of the plume at the northern site was begun in 1991 by means of moored temperature sensors and current meters [Baker and Cannon, 1993].

Fig. 5. (A) Cross-section of EP86 and underlying chronic plume (reprinted from Baker et al. [1989]). Contours are isolines of hydrothermal temperature anomaly (°C). Potential density (kg m-3) isolines shown as dashed lines. Thin line in sawtooth pattern is track of the CTD tow-yo path. (B) Plan view showing locations of EP86 and EP87 in relation to known hydrothermal fields (stars) and volcanic eruption mounds emplaced between 1981 and 1987 (solid beads). Megaplume contours in temperature anomaly. Origin of EP87 is unknown.

Fig. 6. (A) Cross-sections of two event plumes mapped over the CoAxial segment in July, 1993 [Baker et al., 1995]. Both plumes were smaller and had lower temperature anomalies than EP86 or EP87. (B) Plan view of all three CoAxial event plumes showing their relative locations when first discovered. EP93A and EP93B were found directly over the lava eruption mound (solid bead) and warm-water fissure (heavy line) near 46°30'N. EP93C was found two weeks after EP93A and EP93B; its discharge site is unknown.

The Cleft segment time series is proving especially useful in revealing the linkage between hydrothermal and magmatic activity, because it began simultaneously with the observation of a new class of plume phenomena: "event plumes." In 1986, Baker et al. [1987b] discovered a "megaplume" at the northern end of the Cleft segment (Figure 5). Analysis of this and a second megaplume found over the Vance segment in 1987 [Baker et al., 1989; Gendron et al., 1993] demonstrate that event plumes are near-instantaneous releases of enormous volumes of hydrothermal fluid; about108 m3 in the case of the 1986 megaplume. The subsequent discovery of volcanic extrusions emplaced at the north Cleft megaplume site sometime between 1983 and 1987 [Chadwick et al., 1991] left little doubt that event plumes are a signature of the creation or reinvigoration of a vent field by local magmatic activity. This hypothesis was confirmed in 1993 by a series of spectacular observations on the CoAxial segment of the JDFR, just north of Axial Volcano. Between June 26 and July 13, 1993, NOAA/VENTS monitoring of the U.S. Navy's SOSUS hydrophone network detected seismic signals from the JDFR between 46°14'N, 129°50'W and 46°31'N, 129°36'W [Fox, in press]. A series of cruises to this area over the next four months found new volcanic flows, apparently new vent fields, and several examples of event plumes with volumes up to 30% of the 1986 megaplume (Figure 6) [Baker et al., 1995; Embley et al., 1995].

Fig. 7. Summary of hydrothermal temperature anomalies observed along the crest of the Juan de Fuca Ridge, including Axial Volcano, from 1985 to 1989. For clarity, segments and the overlying plumes are displayed individually, without overlap. Triangles show location of known venting sites. Anomalies at the northern end of the Cobb segment, 47.4°-47.6°N, are believed to mark an intrusion of geothermally heated bottom water from Cascadia Basin to the east, rather than axial hydrothermal heating (reprinted from Baker and Hammond [1992]). Significant hydrothermal plumes ( 0.02°C) cover ~25% of the Juan de Fuca Ridge. This figure does not reflect recent work at the N. Axial (CoAxial) segment.

In addition to intensive work along the Cleft segment and at other specific sites, the NOAA/VENTS Program also undertook to survey, at least once, as much of the entire JDFR as feasible. As of 1989, about 90% of the ridge from the Cleft to Endeavour segments had been surveyed at least once, and more than 50% at least twice, by a continuous tow-yo transect. The goal of this program was to test the hypothesis that the distribution and intensity of hydrothermal activity are directly related to the variable magmatic budget along the ridge. Baker and Hammond [1992] reported that hydrothermal discharge is strongest on those segments, or portions of segments, where the apparent magmatic budget is highest, as indicated by the degree of along-axis inflation or other morphological characteristics (Figure 7).

Fig. 8. Location map for East Pacific Rise spreading centers. Solid circles indicate locations of axial discontinuities.

Three areas on the EPR, one medium (Gulf of California, 50-60 mm yr-1), one fast (9°-13°N, 100-120 mm yr-1), and one superfast (14°-23°S,155 mm yr-1) spreading, have been surveyed extensively enough to produce maps of plume distributions at multisegment scales (Figure 8). Two of the sites (the fast and superfast regions) have hosted multichannel seismic investigations of the axial crest [Detrick et al., 1987; 1993], so the extent and depth of the axial magma chamber (AMC) furnishes some indication of the apparent magmatic budget along these ridge crest sections.

Fig. 9. A (3He) section through the Gulf of California showing evidence of hydrothermal venting in the Guaymas Basin (reprinted from Lupton [1979]). Triangle shows location of known venting sites in the Guaymas Basin.

The Gulf of California is a series of en echelon spreading centers opening at 50-60 mm yr-1 and creating a string of deep, semi-enclosed basins (Figure 9). Helium isotope profiles in each of the six major basins [Lupton, 1979] imply that only Guaymas Basin, with (3He) values of 65-70% ((3He) = ((R/Ratm) - 1) × 100, where R = 3He/4He in the fluid or atmosphere, i.e., the 3He/4He ratio was 65-70% higher than in atmospheric helium), sustains active venting (Figure 9).

Plumes over the fast-spreading section of the EPR between 8° and 13°N have been systematically studied since at least 1983. Crane et al. [1988] used the same techniques employed by Crane et al. [1985] on the JDFR to identify perturbations in the temperature field between 8°20' and 13°10'N along the ridge crest, but the results were similarly compromised by the equipment used. This same area has since been resurveyed in detail, by Dynamic Hydrocast and CTD casts from 12°08' to 13°12'N in 1986 [Bougault et al., 1990; Charlou et al., 1991b], and by CTD/transmissometer tows and casts from 8°40' to 11°50'N in 1991 [Lupton et al., 1993; Baker et al., 1994; Feely et al., 1994; Mottl et al., 1995]. These studies precisely mapped the distribution of hydrothermal plumes along the ridge crest, revealing a fine-scale agreement between hydrothermal activity and the apparent magmatic budget, as inferred from ridge morphology and subsurface structure [Bougault et al., 1990; Baker et al., 1994].

Fig. 10. Transects of methane and dissolved manganese along the East Pacific Rise determined from three Dynamic Hydrocast tows [after Bougault et al., 1990]. Connected triangles show area of known venting sites.

Fig. 11. Transects of light-attenuation anomaly (continuous tow-yo) [Baker et al., 1994], dissolved Mn, and CH4 (both from discrete samples) [Mottl et al., 1995] obtained in November 1991 along the East Pacific Rise south and north of the Clipperton Transform Zone. Contour intervals are 0.004 m-1 (with an additional dotted contour at 0.002 m-1) for light-attenuation anomaly, 5 nmol/L for Mn, 5 nmol/L for CH4 south of the Clipperton Transform Zone, and 1 nmol/L for CH4 north of the Clipperton Transform Zone. Triangles show location of known venting sites; double-headed arrow shows extent of April 1991 lava eruption [Haymon et al., 1991, 1993]. Note strong CH4 and relatively weak Mn anomalies directly above the eruption zone, in contrast to the reverse trend over presumably "mature" vent fields near 11°18'N. Significant hydrothermal plumes (light-attenuation anomaly >0.004 m-1 or Mn > 10 nmol/L) cover ~38% of the EPR crest surveyed in Figures 10 and 11.

Between 12°08' and 13°12'N, Bougault et al. [1990] and Charlou et al. [1991b] found plumes with dissolved Mn >10 nmol kg-1and CH4 >2.5 nmol kg-1 from 12°24' to 12°53'N, plus a small zone of high values from 12°11' to 12°15'N; plume maxima were centered between 12°40' and 12°50'N (Figure 10). Between 8°40' and 11°50'N, plumes were mapped from distributions of temperature and light-attenuation anomalies [Baker et al., 1994], dissolved Mn and CH4 [Mottl et al., 1995], and particulate chemistry [Feely et al., 1994]. Based on temperature and light-attenuation anomalies from continuous tow-yo data, intense and spatially continuous plumes were found from 8°48' to 8°58'N, 9°29' to 10°01'N and 11°05' to 11°27'N. Chemical data are available only at discrete locations, but the plume maps based on chemical species such as CH4 and dissolved Mn agree with the hydrographic and optical data (Figure 11).

Plume sampling on the southern EPR has been concentrated along the superfast-spreading (150-156 mm yr-1) portion of the EPR between the Garret Transform Fault at 13°30'S and the Easter microplate at ~23°S. The first unequivocal indicator of active hydrothermal venting there was a westward trending 3He plume mapped along a cross-axis transect at 15°S [Lupton and Craig, 1981]. A series of five along-axis hydrocasts between 19°24' and 22°15'S found concentrations of total dissolvable Mn as high as 49 nmol kg-1, with the highest values occurring over an axial high at 21°22'S [Klinkhammer and Hudson, 1986]. More recent and detailed work in this same area (21°20' to 22°40'S) revealed intense plume anomalies in light attenuation near 21°40'S [Krasnov et al., 1992].

Fig. 12. Tow-yo transects of light attenuation along the East Pacific Rise just south of the Garrett Transform Zone [Baker and Urabe, 1994]. Nephelometer data are combined from five tow-yos, comprising >1000 vertical profiles. Contour interval is 0.01 volts, 0.02 m-1 as measured by a simultaneously recording beam transmissometer. Plume signals are present almost everywhere, but increase sharply south of 17°30'S. Significant hydrothermal plumes (nephelometer volts >0.022) cover ~60% of this survey area.

The first large-scale plume survey in this area was completed in 1993 by the Japanese-NOAA Ridgeflux Expedition using detailed chemical sampling and nearly continuous CTD/SUAVE/optical tow-yos between 13°50' and 18°40'S [Urabe et al., 1994; Baker and Urabe, 1994; Massoth et al., 1994; Ishibashi et al., 1994; Feely et al., 1994; Lupton et al., 1994]. Preliminary analyses show that hydrothermal plumes covered ~60% of the entire study area (Figure 12). Plumes were most common south of ~16°30'S, becoming virtually continuous between 17°20' and 18°40'S. The southern portion of the study area was also characterized by CH4/Mn ratios as high as 3.9, particulate S/Fe ratios >1, and Fe/Mn ratios >8. Similarly high CH4/Mn and S/Fe ratios were found in 1991 in plumes at 9°50'N on the EPR [Lupton et al., 1993], 6 months after a seafloor lava eruption at that site [Haymon et al., 1993]. An increased release of heat and volatiles is typical of hydrothermal systems perturbed by a magmatic intrusion [Baker, 1995].

Western Pacific Marginal Basins

The western boundary of the Pacific Ocean is a broad expanse of marginal basins and troughs fragmented by trench-arc systems that mark subduction zones of several tectonic plates [Taylor and Karner, 1983] (Figure 13). Many of these basins contain active spreading centers that likely host hydrothermal activity. Systematic though limited plume sampling has been conducted in the Harve Trough, Lau Basin, North Fiji Basin, Woodlark Basin (Solomon Sea), Manus Basin (Bismark Sea), Mariana Trough, and Okinawa Trough.

Fig. 13. Location map for western Pacific marginal basins. Thin lines are fracture zones, thick lines are spreading centers, and triangled lines are convergent plate edges. 1, Harve Trough; 2, Lau Basin/Valu Fa Ridge; 3, North Fiji Basin; 4, Woodlark Basin; 5, Manus Basin; 6, Mariana Trough; 7, Okinawa Trough.

The 1986 Papatua Expedition [Craig and Poreda, 1987] occupied four hydrocast stations in the Harve Trough, five in the Lau Basin, and three in the Woodlark Basin. Only weak to negligible hydrothermal indications were found in each basin. Subsequent seafloor explorations in the Lau [Fouquet et al., 1991] and Woodlark [Binns et al., 1993] basins, however, have discovered active hydrothermal sites, indicating that sampling during the Papatua Expedition was too sparse to adequately survey these large basins.

Fig. 14. Transects of methane and dissolved Mn along the central spreading axis in the North Fiji Basin. Data compiled from published and unpublished sources (see text) spanning several years. Note that the data base is not identical for each tracer. Contour intervals are 1 nmol kg-1 (with additional dotted contour at 0.5 nmol kg-1) for CH4, and10 nmol kg-1 (with addition dotted contour at 5 nmol kg-1) for Mn. Triangle shows location of known venting sites. Insets show geographic location of profiles. The distributions suggest other venting sites north and south of the triple junction at 17°S.

Plumes in the North Fiji Basin have been sampled more intensively than in any other back-arc basin. Geophysical surveys [Auzende et al., 1994] indicate two north-south spreading axes: a medium-rate (50-60 mm yr-1) ridge running along ~173°30'E and a slow-spreading ridge along ~176°E. Only the westerly ridge has been sampled. Between 21°S and a triple junction at 16°50'S the morphology is typical of a fast-spreading ridge, shallowing from an axial depth of ~2800 m south of 18°S to <2500 m at the ridge-ridge-transform triple junction. North of the triple junction the ridge trends NNW and assumes a slow-spreading ridge morphology of 3500-4000-m-deep grabens. South of the triple junction, CH4 and dissolved Mn were first sampled on SEAPSO 3 in 1985 [Auzende et al., 1988], again in 1986 (CH4 only) by the Papatua Expedition [Craig and Poreda, 1987; Craig et al., 1987], and in 1987 on Kaiyo 87 [Nojiri et al., 1989] (J. Ishibashi, unpublished data, 1994). Sedwick et al. [1990] sampled total dissolvable Mn north of the triple junction in 1987. An along-axis distribution of the reported values (Figure 14) shows several isolated maxima between 19° and 17°S. The particularly intense CH4 plume at 18°8'S also carried a (3He) maximum of 71%, confirming its magmatic origin [Craig et al., 1987]. Nojiri et al. [1989] interpreted as an event plume a strong plume 600 m above bottom at 18°48'S with a high CH4/Mn ratio that was observed in 1987 but not in 1988. Sedwick et al. [1990] proposed that intense Mn and weak CH4 signals high above the ridge north of the triple junction were advective features from venting near the triple junction [Grimaud et al., 1991], attributing the absence of CH4 to in situ consumption or a deficit of volatiles in the original discharge (e.g., hydrothermal brine from a phase-separated system).

Fig. 15. Transects of methane and manganese collected in 1990 along the spreading center in the eastern Manus Basin (reprinted from Gamo et al. [1993] with kind permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington OX5 16B, UK). The Papatua Expedition [Craig and Poreda, 1987] collected samples near stations 37 and 29. Triangle shows location of known venting sites.

The Manus Basin is a fast-spreading (>100 mm yr-1) basin with two east-northeast trending spreading axes connected by a northwest trending transform fault (Figure 13) [Taylor, 1979]. The Papatua Expedition occupied three stations over the northern ridge, finding CH4 anomalies up to 3.5 nmol kg-1 and a (3He) maximum of 58%. The Papatua Expedition also found CH4 anomalies over the southern ridge, which was sampled more extensively by Gamo et al. [1993] in 1990. Gamo et al. [1993] mapped multiply layered CH4 and Mn plumes between 1000 and 2000 m depth that indicated at least two distinct hydrothermal sites (Figure 15). CH4 concentrations in the deep plume remained steady at 1-2 nmol kg-1 between the 1986 Papatua Expedition and 1990, but no indication of shallower plumes was seen in 1986. This appearance of a shallow plume and its high CH4/Mn ratio led Gamo et al. [1993] to suggest that it was an event plume. The extremely small thickness/width ratio (<0.004) of the shallow plume relative to known event plumes (~0.03 [Baker, 1994]), however, argues against this hypothesis.

Farther to the north, plumes have been mapped in the Okinawa and Mariana Troughs. In the Okinawa Trough, a relatively shallow basin behind the deep Ryukyu Trench, strong CH4, Mn, and 3He (up to a (3He) of 65%) anomalies at depths between 1300 and 1600 m were found in 1987 [Ishibashi et al., 1988] and 1988 (J. Ishibashi, personal communication, 1994). Horibe et al. [1986] occupied a series of hydrocasts between 18°11'N and 18°15'N in 1982 in the Mariana Trough, mapping CH4 plumes near-bottom and 700-800 m above the seafloor. A resurvey of this same area in 1986 [Horibe and Craig, 1987] again found CH4 plumes, though somewhat differently distributed. The CH4 plumes in both years had no 3He signature.

Mid-Atlantic Ridge

Unlike fast-spreading ridge axes such as the EPR, the neovolcanic zone of the slow-spreading MAR (Figure 16) sits not at a narrow axial high, but within a broad (5-10 km wide) axial rift valley. The typical valley is bounded across axis by steep and faulted valley walls that rise 1000-2000 m above the valley floor and along axis by frequent transform-fault offsets. The height of rise of buoyant hydrothermal plumes is nearly always less than the bounding heights of the MAR rift valley. Consequently, hydrothermal plumes are trapped and transported within the axial valley. These bathymetric characteristics allow easy detection of hydrothermal tracers within a particular segment, but make precise location and enumeration of the discharge sources difficult [e.g., Klinkhammer et al., 1985; Nelsen et al., 1986/87; Charlou and Donval, 1993; Murton et al., 1994]. Plume mapping on the MAR has typically covered a greater linear distance at significantly less detail than similar studies in the eastern Pacific.

Fig. 16. Location map for the northern Mid-Atlantic Ridge and Reykjanes Ridge. Known and suspected hydrothermal sites identified in Figure 19.

Fig. 17. Along-axis maxima in total reactive (~dissolved) Mn [Klinkhammer et al., 1985] and CH4 [Charlou and Donval, 1993] from vertical casts on the Mid-Atlantic Ridge. High values in parentheses at 26°N and 29°N are representative samples over the TAG [Klinkhammer et al., 1986] and Broken Spur [Elderfield et al., 1993] hydrothermal fields, respectively. Thin dashed line on each plot marks the regional background value of Mn and CH4.

Until 1984 it was widely predicted that hydrothermal activity might be restricted to fast-spreading ridges and that at slow-spreading ridges, such as the MAR, heat flow would be insufficient to support active high-temperature black smoker vent fields. It is now clear that this is not the case. In 1984, Klinkhammer et al. [1985] carried out systematic discrete sampling and shipboard Mn analysis of seawater samples from stations occupied throughout the MAR rift valley between 11°N (the Vema Fracture Zone) and 26°N (Figure 17). Nine discrete segments were occupied during this survey and dissolved Mn anomalies--interpreted as the product of active hydrothermal venting--were found at every station. As a conservative estimate, Klinkhammer et al. [1985] concluded that a minimum of at least five discrete hydrothermal fields must characterize this section of the MAR.

Fig. 18. Geochemical plume transect for total reactive Mn at the TAG hydrothermal site in 1985 (reprinted from Klinkhammer et al. [1986] with kind permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kiddington 0X5 16B, UK). Solid circles indicate sample positions.

These predictions were confirmed in 1985 when a series of CTD-nephelometer hydrocasts (Figure 18) and deep-tow video camera deployments confirmed the first active black-smoker hydrothermal field at the northernmost station occupied by Klinkhammer et al. [1985], the TAG hydrothermal field at 26°N [Rona et al., 1986]. In 1986, black-smoker hydrothermal venting was also found at the Snakepit hydrothermal field at 23°N [Ocean Drilling Program Leg 106 Scientific Party, 1986], close to where Klinkhammer et al. [1985] had reported water column dissolved Mn anomalies.

Between 1985 and 1988 another series of cruises mapped the distribution of CH4 between 12° and 26°N [Charlou and Donval, 1993] (Figure 17). High CH4 concentrations were found over the TAG and Snakepit sites, as well as the 15°20' fracture zone. Because of very low Mn in the plumes near the 15°20' fracture zone, and the occurrence of local outcrops of serpentinized ultrabasic diapirs, Charlou and Donval [1993] hypothesized that CH4 anomalies there arise from fluid circulation in ultrabasic rocks rather than basalt-seawater interactions typical of faster-spreading ridges. The changes of mechanical properties and density caused by the serpentinization of deep crustal rocks, as inferred from these CH4 anomalies, may play an influential role in the construction of slow-spreading ridges [Charlou et al., 1991a; Charlou and Donval, 1993; Bougault et al., 1993].

Fig. 19. Summary of evidence for hydrothermal venting along the Mid-Atlantic Ridge compiled from water-column investigations between 11° and 40°N [German et al., 1995]. Only four segments or segment portions have confirmed active venting.

In 1992, as part of a joint FARA (French American Ridge Atlantic) cruise aboard the R/V Atlantis II, Project FAZAR, systematic water-column sampling was carried out along a more northerly section of the Mid-Atlantic Ridge between 32°N and the Kurchatov Fracture Zone close to 40°N [e.g., Wilson et al., in press; Klinkhammer et al., in press]. Stations were occupied approximately every 15 miles along axis using a CTD/nephelometer/transmissometer array augmented with the ZAPS Mn probe. Eleven of 19 segments were investigated and evidence for high-temperature black-smoker-type hydrothermal activity was found in seven, including the AMAR segment at 36°N and the confirmed Lucky Strike hydrothermal field at 37°17'N [Langmuir et al., 1993; Wilson et al., in press; Klinkhammer et al., in press]. Further studies south of the Atlantis fracture zone [Murton et al., 1994; Chin et al., 1994] have used a combination of a transmissometer and the ZAPS sensor mounted on IOS Deacon Laboratory's Towed Ocean Bottom Instrument (TOBI) to identify discrete hydrothermal signals in three further segments, between 27°N and 30°N, including the Broken Spur hydrothermal field at 29°N. Thus, systematic water-column surveys have been carried out along most of the Mid-Atlantic Ridge rift valley between 11°N and 40°N and evidence for at least 15 sites of high-temperature hydrothermal venting has been found. This yields an average incidence of hydrothermal activity of at least one site every ~175 km for the ~2,500 km of ridge-crest surveyed (Figure 19) [German et al., 1995].

Reykjanes Ridge

German et al. [1994] carried out a systematic survey of the Reykjanes Ridge between 57°45'N and 63°09'N. Approximately 120 stations were occupied at 10-20 km intervals (i.e., at closer spacings than the 11-26°N and 32-40°N surveys), using a CTD/nephelometer/transmissometer array combined with shipboard analyses for total dissolvable Mn and dissolved Si, CH4, and H2. The only evidence of hydrothermal activity throughout this 750 km of ridge crest was found at the Steinahóll vent field at 63°06'N. The site is situated in just 250-350 m of water and is notable for the formation of bubble-rich plumes that have been imaged using a high-frequency (38 kHz) echo sounder (Figure 20). It is important to understand why such a lengthy section of ridge, patently affected by the local thermal flux of the Icelandic hot-spot, should appear to be so devoid of high-temperature hydrothermal activity.

Fig. 20. 38 kHz echo-sounder trace from a transect along the axis of the Reykjanes Ridge between 63°06.04'N and 63°06.10'N, indicating the presence of gas-rich hydrothermal plumes rising from the seabed close to 63°06.06'N (reprinted from German et al. [1994] with kind permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington OX5 1GB, UK).

Indian Ocean

Very few studies of hydrothermal activity have been carried out in the Indian Ocean, even though sediments exhibiting hydrothermal metal enrichments along the Indian Ocean ridge system have been known for almost thirty years [Boström et al., 1969]. To date, however, discovery of evidence for hydrothermal activity along the ridges of the Indian Ocean has been limited to three specific studies. The Central Indian Ridge between 21°S and 24°S, close to the Rodriguez Triple Junction, was investigated between 1983 and 1988 by cruises of the R/V Sonne. A series of CTD hydrocast stations occupied between 21°15'S and 21°30'S in 1986 found distinct plumes of dissolved Mn and CH4 anomalies at several stations, with concentrations up to 27.5 nmol kg-1 and 2 nmol l-1 respectively [Herzig and Plüger, 1988]. Evidence for an additional site of hydrothermal activity was revealed in 1988 by the presence of ~20 nmol kg-1 dissolved Mn anomalies and up to 8.9 nmol l-1 CH4 anomalies at ~24°S [Plüger et al., 1990].

In 1993, a group of Japanese investigators conducted water-column observations around the Rodriguez Triple Junction [Gamo et al., 1994]. Hydrothermal optical and chemical (CH4 and dissolved Mn and Fe) anomalies were found centered around the 2250 m horizon near 25°20'N over the eastern rift wall of the Central Indian Rift, the northern arm of the Rodriguez Triple Junction. Because of the shallow depth of these plumes and the lack of strong anomalies directly over the 4000-m deep rift valley axis, the source is likely located on the eastern rift valley wall [Gamo et al., 1994]. Two casts at the 24°N site of the plume anomalies discovered by Plüger et al. [1990] found CH4 and dissolved Mn anomalies similar to the 1988 profiles. Only negligible anomalies were found on casts in the Southeastern and Southwestern Indian Ridge valleys, the other two arms of the Rodriguez Triple Junction. Based on the distribution of deep 3He anomalies in the Indian Ocean, the Rodriguez Triple Junction area appears to be the largest source of hydrothermal effluent from Indian Ocean ridges [Jamous et al., 1992].

DISCUSSION

Exploration Implications

The studies described above document that about 10% of the 60,000 km of global spreading center systems have been examined at least cursorily for chemical or physical evidence of hydrothermal plumes. Perhaps one-third of that fraction has been mapped in enough detail to confidently locate all the major vent fields on a particular segment (e.g., Figure 7, 10, 11, 12, 15). The principal result of these systematic explorations is that hydrothermal venting is present, albeit at widely differing levels of activity, across the entire spectrum of ridge crest spreading rate, topographic expression, magmatic budget, and other geological and geophysical characteristics. Many other generalizations have crystallized from plume studies over the last decade and are discussed herein and in other papers within this volume [e.g., Lupton, this volume; Lilley et al., this volume; Kadko et al., this volume; McDuff, this volume], but here we wish to emphasize an important operational result that remains underappreciated by the ridge-crest community at large: the utility of plumes as exploration tools for seafloor vents.

In the heady early days of hydrothermal discovery, many workers believed that the chemistry of hydrothermal solutions was simply a function of exit temperature, and that this function could be established empirically by careful sampling at a few sites [e.g., Edmond et al., 1982]. Further discoveries demonstrated a much more complex universe of chemical variations [e.g., Von Damm, 1990], emphasizing the need for continued exploration and the compilation of as large a hydrothermal data base as possible. (We leave unwritten the many biological, physical, and geological reasons for vent field exploration.)

The search for new vent fields has historically followed an inefficient path: A tectonic segment, or segment portion, is chosen on geological considerations. A remote or manned search for geologic or biologic visual clues of active venting is then mounted to pinpoint discharge locations. Finally, vent fluids are sampled from a submarine and plume waters from a surface ship. Consideration of the several studies reviewed in this paper demonstrates that a more efficient approach would postpone detailed seafloor exploration until after systematic plume mapping had revealed the relative magnitude and approximate location of discharge sources. Moreover, chemical and physical observations of plume waters can provide first-order information on the composition of the source fluids, which are likely indicative of the present magmatic and hydrothermal conditions of the vent field.

The efficacy of plume reconnaissance can be documented from several recent examples. On the EPR near 10°N, an expensive and exhaustive (~107 m2) video survey by the Argo imaging system between 9°09' and 9°54'N in 1989 [Haymon et al., 1991], followed by a submersible program in 1991 [Haymon et al., 1993], located virtually all discharge sources along the surveyed ridge crest. A schematic map of these sources compared with the plume distribution from a chemical and hydrographic water-column survey [Lupton et al., 1993; Baker et al., 1994] shows excellent agreement, at roughly the kilometer scale, even for isolated single smokers (Figure 11). Various plume tracers, especially CH4 and Mn [Mottl et al., 1995], identify the location and chemical character of unusual fluids rich in volatiles (e.g., CH4, 3He, H2S) and low in chlorinity (and thus dissolved metals) associated with a recent dike intrusion and lava eruption at 9°54'N [Haymon et al., 1993]. Plume measurements can thus provide first-order information on the magmatic and hydrothermal history of the vent field independent of seafloor sampling.

While the above example is an after-the-fact plume success story, instances of plume mapping directing the discovery of seafloor discharge sources are common. On the CoAxial segment of the JDFR, plume surveying immediately after the remote detection of seismic activity in 1993 directed seafloor imagery mapping by a remotely operated vehicle towards the eventual discovery and sampling of a new lava eruption [Embley et al., 1995]. On the MAR, systematic plume surveying between 27° and 30°N in 1993 discovered optical and chemical evidence for three possible hydrothermal sites, including a strong signal at 29°10.1'N, 43°10.3'W [Murton et al., 1994]. The Broken Spur vent field was confirmed at this site three months later during an Alvin dive series [Murton et al., 1994].

Fig. 21. Comparison of the global extent of plume studies and seafloor vent fluid sampling. Shaded areas on spreading centers mark the locations of multi-segment plume mapping investigations on the Reykjanes Ridge, Mid-Atlantic Ridge, northeast Pacific ridges, Gulf of California, northern and southern East Pacific Rise, and the north Fiji Basin. Numbers refer to Table 1 and mark sites of vent fluid collection and analysis along spreading centers; fluid discharge sites noted only from visual observations are not included.

The number of fluid discharge sites sampled by submersible or remote vehicle is steadily growing but remains only a tiny fraction of all global sources (Table 1, Figure 21). Plume mapping and sampling has allowed us to expand our knowledge of the distribution and geologic setting of hydrothermal venting beyond the limits imposed by camera and submersible operations. Figure 21 graphically compares the relative coverage of vent and plume data sets. The expansion of hydrothermal information afforded by plume studies is essential for regional-scale issues such as the effect of hydrothermal cooling on heat transport within the crust [Phipps Morgan and Chen, 1993], the movement of melt within the crust [Purdy et al., 1992], magmatic differentiation [Sinton et al., 1991], and the effect of hydrothermal heating on abyssal waters [Hautala and Riser, 1993].

If your browser cannot view the following table correctly, click this link for a GIF image of Table 1.

TABLE 1. Sites of Hydrothermal Vent Fluid Collection


Map No. Site Reference

1 Lucky Strike (37°17'N) Colodner et al. [1993]
2 Broken Spur (29°N) James et al. [in press b]
3 TAG (26°N) Campbell et al. [1988]
4 Snakepit/MARK (23°N) Campbell et al. [1988]
5 Explorer Ridge (49°45'N) Tunnicliffe et al. [1986]
6 Middle Valley Butterfield et al. [1994b]
7 Endeavour (48°) Butterfield et al. [1994a]
8 Axial Volcano (46°N) CASM [1985]
9 Cleft--North (45°) Butterfield and Massoth [1994]
10 Cleft--South (44°40'N) Von Damm and Bischoff [1987]
11 Escanaba Trough Campbell et al. [1994]
12 Guaymas Basin (27°N) Von Damm et al. [1985a]
13 EPR (21°N) Von Damm et al. [1985b]
14 EPR (13°N) Michard et al. [1984]
15 EPR (11°N) Bowers et al. [1988]
16 EPR (10°N) Von Damm et al. [1991]
17 Galapagos (86°W) Edmond et al. [1979]
18 EPR (17°-19°S) Charlou et al. [1994]
19 Okinawa Trough (27°30'N) Sakai et al. [1990]
20 Mariana Trough Campbell et al. [1987]
21 Manus Basin Lisitsyn et al. [1993]
22 Woodlark Basin Binns et al. [1993]
23 N. Fiji Basin (17°S) Grimaud et al. [1991]
24 Lau/Valu Fa (22°S) Fouquet et al. [1991]

Geological Implications

As an example of the kind of linkage between geological and hydrothermal data made possible by plume studies, we examine a simple but crucial hypothesis: A functional relationship exists between axial hydrothermal heat flux and ridge-crest spreading rate, a convenient analogue of the time-averaged magmatic budget. Incorporation of hydrothermal cooling into models of the genesis of oceanic crust [e.g., Phipps Morgan and Chen, 1993] requires generalizations of this kind. To use plume distributions to test this hypothesis, we assume that hydrothermal heat output is proportional to plume incidence, the percentage of ridge-axis length overlain by a significant hydrothermal plume. Simply put, more hydrothermal cooling produces more plumes. (We operationally define a "significant" plume as one with plume tracer values greater than three standard deviations above the background (non-plume) mean.) Our assumption of the equivalence of heat flux and plume incidence is reasonable over the narrow crest and shallow axial depressions characteristic of ridges spreading faster than ~60 mm yr-1 [Baker and Hammond, 1992; Baker, 1994; Baker et al., 1994]. On these ridges, plumes are advected off axis quickly and plume incidence closely matches the extent of the vent fields. On slow-spreading ridges with deep axial valleys, the high bounding walls trap plumes so water over an entire segment can become "contaminated" with hydrothermal anomalies.

Fig. 22. (a) Full-rate seafloor spreading vs. axial magma chamber depth from seismic observations [Purdy et al., 1992] (solid circles; LAU, Valu Fa Ridge in the eastern Lau Basin; JDFR, Cobb segment on the Juan de Fuca Ridge; NEPR, northern East Pacific Rise, 9°-13°N; SEPR, southern EPR, 14°-20°S) and from the Phipps Morgan and Chen [1993] crustal thermal model (solid line). (b) Full-rate seafloor spreading vs. plume incidence (solid squares) and hydrothermal cooling in the neovolcanic zone according to a crustal thermal model (Y. J. Chen and J. Phipps Morgan, The effect of magma emplacement geometry, spreading rate, and crustal thickness on hydrothermal heat flux at mid-ocean ridges, submitted to Geophys. Res. Lett., 1994) (dashed line). Plume incidence data available for the Reykjanes Ridge (RR) [German et al., 1995], Mid-Atlantic Ridge (MAR) (11°-40°N) [German et al., 1994], JDFR [Baker and Hammond, 1992], NEPR (8°40-13°10'N) [Bougault et al., 1990; Baker et al., 1994], and SEPR (13°50'-18°40'S) [Baker and Urabe, 1994]. "Plume incidence" is the percentage of axis length overlain by a significant (plume tracer values greater than background mean plus three standard deviations) plume. The heavy solid line is a least-squares fit forced through the origin: slope = 0.375, r2 = 0.98. Hydrothermal cooling is the steady state convective heat flux from a neovolcanic zone 2 km wide centered on the spreading axis.

Continuous surveys from the JDFR (Figure 7) [Baker and Hammond, 1992], the northern EPR (Figures 10, 11) [Bougault et al., 1990; Baker et al., 1994], and the southern EPR (Figure 12) [Baker and Urabe, 1994] have yielded direct measures of plume incidence. On slow-spreading ridges, in the current absence of conclusive continuous surveys, we approximate plume incidence from estimates of vent field frequency derived from discontinuous plume surveys along the MAR and Reykjanes Ridge [German et al., 1994; German et al., 1995]. If we assign a reasonable plume size of 10 km to each of the 15 suspected sites of venting on the MAR between 11° and 40°N [German et al., 1995], we calculate a 6% plume incidence. Similarly, the one vent field found on the Reykjanes Ridge [German et al., 1994] yields a 2% plume incidence. When data from all five regions are plotted together, a strong linear relationship between plume incidence and spreading rate emerges (Figure 22).

It is vital to realize that Figure 22 is valid only over long time or space scales. Vigorous hydrothermal activity can occur on segments of any spreading rate, so an instantaneous observation at a particular location may not be representative of its long-term hydrothermal history. We obtain a representative hydrothermal history of a particular ridge-crest location only by observing it through geologic time, which is impractical, or by simultaneously observing many individual segments spreading at the same rate. Figure 22 thus implies that, over long time scales, (1) one-third of the global ridge axis spreading at ~100 mm yr-1 (for example) will be hydrothermally active at any instant, and (2) any area on that ridge-axis portion will be hydrothermally active one-third of the time.

How does the plume vs. spreading rate relationship compare to other observed and modeled parameters? Axial magma chamber (AMC) depth from seismic observations [Purdy et al., 1992] shows a rapid decrease from 60 to 120 mm yr-1 followed by only a slight further shallowing at superfast rates. This same trend has been reproduced by the crustal thermal model of Phipps Morgan and Chen [1993], in which the supply of heat by magma injection and the removal of heat by hydrothermal cooling is balanced to produce a steady state magma lens at crustal depths corresponding to the seismic observations at various spreading rates (Figure 22). In contrast to the non-linear relationship between magma chamber depth and spreading rate, the relationship between spreading rate and the hydrothermal cooling required by the crustal thermal model (Y. J. Chen and J. Phipps Morgan, The effect of magma emplacement geometry, spreading rate, and crustal thickness on hydrothermal heat flux at mid-ocean ridges, submitted to Geophys. Res. Lett., 1994) is approximately linear (Figure 22). The common linear trend between spreading rate and both plume incidence and hydrothermal cooling supports our initial assumption that plume incidence is roughly proportional to hydrothermal heat flux from the neovolcanic zone of ridge crests.

Still at issue is whether our meager coverage of the global ridge crest is representative and the trends described above are general. Over 3900 km of the southern EPR spreads faster than 140 mm yr-1 [DeMets et al., 1990], and we have a single snapshot of hydrothermal activity along only an eighth of that distance. Other spreading-rate portions of the midocean ridge crest are similarly undersampled, especially the slow-spreading ridges. A confident understanding of the balance between global (e.g., spreading rate) and local (e.g., magma supply, crustal permeability) controls on the distribution and vigor of hydrothermal activity awaits continued exploration.

SUMMARY

Hydrothermal plumes formed by mixing of seafloor vent fluids and ambient seawater are easily detectable by physical and chemical tracers. New in situ instrumentation is providing the capability of continuous measurement of certain chemical species (e.g., dissolved Mn and Fe) to complement the continuous measurement of hydrographic and optical tracers. Hydrothermal plume evidence of seafloor venting has been recognized for over 30 years, though only in the last decade have systematic surveys been conducted. Most detailed surveys have been conducted along medium- to superfast-spreading eastern Pacific spreading centers because the narrow crest of these ridges permits a largely two-dimensional mapping effort. Carefully studied areas include the Juan de Fuca Ridge (44°30'-48°30'N) and 9°-13°N and 13°40'-18°40'S on the East Pacific Rise. Plumes have been detected in several western Pacific marginal basins, but only parts of the North Fiji (15°-20°S) and Manus Basins have been surveyed in detail. In the North Atlantic, plume surveys have visited a longer length of ridge than in the Pacific, but surveys typically have been less detailed because of the wide and deep axial valleys. Work along the Reykjanes Ridge (58°-63°N) and the Mid-Atlantic Ridge (11°-40°N) indicates a lower incidence of venting than in the Pacific. Plumes have been sampled near the Rodriguez Triple Junction in the Indian Ocean, but no systematic mapping has been accomplished there. About 10% of the 60,000 km of the global spreading center system has been at least cursorily explored for hydrothermal plumes, although only one-third of that total has been mapped in enough detail to confidently locate vent locations.

Two primary conclusions are apparent from this review. First, careful plume surveys can efficiently locate seafloor discharge sites to within a few kilometers or less. Plume mapping can substantially diminish the search time required to locate vents precisely by submersible or remote imagery, and should normally precede seafloor exploration efforts to improve their efficiency. Second, while seafloor venting occurs on ridge segments of spreading rates from slow to superfast, hydrothermal plume production along any multisegment ridge portion is directly related to the spreading rate.

Acknowledgments. This review was supported by the NOAA VENTS Program (ETB), by IOSDL and NERC grant BRIDGE 15 (CRG), and by NERC grant GR3/8596 and a grant from the NERC BRIDGE Program (HE). The impetus for this paper was discussions at the 1993 RIDGE Theoretical Institute at Big Sky, MT, USA. We thank the conveners and participants for channeling our energies into this effort. Cambridge Earth Sciences contribution number ES 4086. IOSDL contribution number 95003. PMEL contribution number 1538.

REFERENCES

Auzende, J. M., J. P. Eissen, Y. Lafoy, P. Gente, and J. L. Charlou, Seafloor spreading in the North Fiji Basin (southwest Pacific), Tectonophysics, 146, 317-351, 1988.

Auzende, J. M., B. Pelletier, and Y. Lafoy, Twin active spreading ridges in the North Fiji Basin (southwest Pacific), Geology, 63-66, 1994.

Baker, E. T., A 6-year time series of hydrothermal plumes over the Cleft segment of the Juan de Fuca Ridge, J. Geophys. Res., 99, 4889-4904, 1994.

Baker, E. T., Characteristics of hydrothermal discharge following a magmatic intrusion, in <i>Hydrothermal Vents and Processes</i>, edited by D. Dixon, L. M. Parson, and C. L. Walker, Geol. Soc. London Spec. Pub. No. 87, 65-76, 1995.

Baker, E. T., and G. A. Cannon, Long-term monitoring of hydrothermal heat flux using moored temperature sensors, Cleft segment, Juan de Fuca Ridge, Geophys. Res. Lett., 20, 1855-1858, 1993.

Baker, E. T., and S. R. Hammond, Hydrothermal venting and the apparent magmatic budget of the Juan de Fuca Ridge, J. Geophys. Res., 97, 3443-3456, 1992.

Baker, E. T., and G. J. Massoth, Hydrothermal plume measurements: A regional perspective, Science, 234, 980-982, 1986.

Baker, E. T., and G. J. Massoth, Characteristics of hydrothermal plumes from two vent fields on the Juan de Fuca Ridge, northeast Pacific Ocean, Earth Planet. Sci. Lett., 85, 59-73, 1987.

Baker, E. T., and T. Urabe, Distribution of hydrothermal plumes along the superfast-spreading East Pacific Rise, 13°50'-18°40'S, Eos Trans. AGU, 75(44 Supplement), 321, 1994.

Baker, E. T., G. J. Massoth, R. W. Collier, J. H. Trefry, D. Kadko, T. A. Nelsen, P. A. Rona, and J. E. Lupton, Evidence for high-temperature hydrothermal venting on the Gorda Ridge, Northeast Pacific Ocean, Deep-Sea Res., 34, 1461-1476, 1987a.

Baker, E. T., G. J. Massoth, and R. A. Feely, Cataclysmic hydrothermal venting on the Juan de Fuca Ridge, Nature, 329, 149-151, 1987b.

Baker, E. T., J. W. Lavelle, R. A. Feely, G. J. Massoth, S. L. Walker, and J. E. Lupton, Episodic venting of hydrothermal fluids from the Juan de Fuca Ridge, J. Geophys. Res., 94, 9237-9250, 1989.

Baker, E. T., R. E. McDuff, and G. J. Massoth, Hydrothermal venting from the summit of a ridge-axis seamount: Axial Volcano, Juan de Fuca Ridge, J. Geophys. Res., 95, 12,843-12,854, 1990.

Baker, E. T., R. A. Feely, M. J. Mottl, F. J. Sansone, C. G. Wheat, J. A. Resing, and J. E. Lupton, Hydrothermal plumes along the East Pacific Rise, 8°40' to 11°50'N: Plume distribution and relationship to the apparent magmatic budget, Earth Planet. Sci. Lett. 128, 1-17, 1994.

Baker, E. T., G. J. Massoth, R. A. Feely, R. W. Embley, R. E. Thomson, and B. J. Burd, Hydrothermal event plumes from the CoAxial seafloor eruption site, Juan de Fuca Ridge, Geophys. Res. Lett., 22(2), 147-150, 1995.

Binns, R. A., S. D. Scott, Y. A. Bogdanov, A. P. Lisitzin, V. V. Gordeev, E. G. Gurvich, E. J. Finlayson, T. Boyd, L. E. Dotter, G. E. Wheller, and K. G. Muravyev, Hydrothermal oxide and gold-rich sulfate deposits of Franklin Seamount, western Woodlark Basin, Papua, New Guinea, Econ. Geol., 88, 2122-2153, 1993.

Bolger, G. W., P. R. Betzer, and V. V. Gordeev, Hydrothermally derived manganese suspended over the Galapagos spreading center, Deep-Sea Res., 25, 721-733, 1978.

Bostrom, K., M. N. A. Peterson, O. Joensuu, and D. E. Fisher, Aluminum-poor ferromangoan sediments on active ocean ridges, J. Geophys. Res., 74, 3261-3270, 1969.

Bougault, H., J. L. Charlou, Y. Fouquet, and H. D. Needham, Activité hydrothermale et structure axiale des dorsales Est-Pacifique et médio-Atlantique, Oceanol. acta vol. special 10, 199-207, 1990.

Bougault, H., J. L. Charlou, Y. Fouquet, H. D. Needham, N. Vaslet, P. Appriou, P. Jean Baptiste, P. A. Rona, L. Dmitriev, and S. Silantiev, Fast and slow spreading ridges: Structure and hydrothermal activity, ultramafic topographic highs and CH4 output, J. Geophys. Res., 98, 9643-9651, 1993.

Bowers, T. S., A. C. Campbell, C. I. Measures, A. J. Spivack, M. Kadem, and J. M. Edmond, Chemical controls on the composition of vent fluids at 13°-11°N and 21°N, East Pacific Rise, J. Geophys. Res., 93, 4522-4536, 1988.

Butterfield, D. A., and G. J. Massoth. Geochemistry of north Cleft segment vent fluids: Temporal changes in chlorinity and their possible relationship to recent volcanism, J. Geophys. Res., 99, 4159-4168, 1994.

Butterfield, D. A., R. E. McDuff, M. J. Mottl, M. D. Lilley, J. E. Lupton, and G. J. Massoth, Gradients in the composition of hydrothermal fluids from the Endeavour segment vent field: Phase separation and brine loss, J. Geophys. Res., 99, 9561-9583, 1994a

Butterfield, D. A., R. E. McDuff, J. M. Franklin, and C. G. Wheat, Geochemistry of hydrothermal vent fluids from Middle Valley, Juan de Fuca Ridge, Chapter 20 in Proceedings of the Ocean Drilling Program, Scientific Results, 139, M.J. Mottl, E.E. Davis, A.T. Fisher, and J.F. Slack (eds.), College Station, TX, 395–410, 1994b.

Campbell, A. C., J. M. Edmond, D. Colodner, M. R. Palmer, and K. K. Falkner, Chemistry of hydrothermal fluids from the Mariana Trough back arc basin in comparison to mid-ocean ridge fluids, Eos Trans. AGU, 68, 1531, 1987.

Campbell, A. C., M. R. Palmer, G. P. Klinkhammer, T. S. Bowers, J. M. Edmond, J. R. Lawrence, J. F. Casey, G. Thompson, S. Humphris, P. Rona, and J. A. Karson, Chemistry of hot springs on the Mid-Atlantic Ridge, Nature, 335, 514-519, 1988.

Campbell, A. C., C. R. German, M. R. Palmer, T. Gamo, and J. M. Edmond, Chemistry of hydrothermal fluids from the Escanaba Trough, Gorda Ridge, in Geologic, Hydrothermal, and Biologic Studies at Escanaba Trough, Gorda Ridge, Offshore Northern California, edited by J. L. Morton, R. A. Zierenberg, and C. A. Reiss, U. S. Geol. Sur. Bull. 2022, Chapter 11, 201-221, 1994.

Cannon, G. A., D. J. Pashinski, and M. R. Lemon, Mid-depth flow near hydrothermal venting sites on the southern Juan de Fuca Ridge, J. Geophys. Res., 96, 12,815-12,831, 1991.

Canadian American Seamount Expedition, Hydrothermal vents on an axis seamount of the Juan de Fuca Ridge, Nature, 313, 212-214, 1985.

Chadwick, Jr., W. W. , R. W. Embley, and C. G. Fox, Evidence for volcanic eruption on the southern Juan de Fuca Ridge between 1981 and 1987, Nature, 313, 212-214, 1991.

Charlou, J. L., and J. P. Donval, Hydrothermal methane venting between 12°N and 26°N along the Mid-Atlantic Ridge, J. Geophys. Res., 98, 9625-9642, 1993.

Charlou, J. L., H. Bougault, P. Appriou, T. Nelsen, and P. Rona, Different TDM/CH4 hydrothermal plume signatures: TAG site at 26°N and serpentinized ultrabasic daipir at 15°05'N on the Mid-Atlantic Ridge, Geochim. Cosmochim. Acta, 55, 3209-3222, 1991a.

Charlou, J. L., H. Bougault, P. Appriou, P. Jean-Baptiste, J. Etoubleau, and A. Birolleau, Water column anomalies associated with hydrothermal activity between 11°40' and 13°N on the East Pacific Rise: discrepancies between tracers, Deep-Sea Res., 38, 569-596, 1991b.

Charlou, J. L., J. P. Donval, Y. Fouquet, and J. M. Auzende, Variability of hydrothermal fluids composition on ultra fast spreading ridge segments between 17° and 19°S on the EPR (NAUDUR cruise), Eos Trans. AGU, 75(44 Supplement), 320, 1994.

Chin, C. S., K. H. Coale, V. A. Elrod, K. S. Johnson, G. J. Massoth, and E. T. Baker, In situ observations of dissolved iron and manganese in hydrothermal vent plumes, Juan de Fuca Ridge, J. Geophys. Res., 99, 4969-4984, 1994.

Clarke, W. B., M. A. Beg, and H. Craig, Excess helium-3 in the sea: evidence for terrestrial primordial helium, Earth Planet. Sci. Lett., 6, 213-220, 1969.

Coale, K. H., C. S. Chin, G. J. Massoth, K. S. Johnson, and E. T. Baker, In situ chemical mapping of dissolved iron and manganese in hydrothermal plumes, Nature, 352, 325-328, 1991.

Collier, R. W., and E. T. Baker, Hydrography and geochemistry of northern Gorda Ridge, in Gorda Ridge, edited by G. R. McMurray, Springer-Verlag, pp. 21-29, 1990.

Colodner D., J. Lin, K. Von Damm, L. Buttermore, R. Kozlowski, J. L. Charlou, J. P. Donval and C. Wilson, Chemistry of the Lucky Strike hydrothermal fluids: initial results, Eos Trans. AGU, 74(43 Supplement), 99, 1993.

Corliss, J. B., J. Dymond, L. I. Gordon, J. M. Edmond, R. P. von Herzen, R. D. Ballard, K. Green, D. Williams, A. Bainbridge, K. Crane, and T. H. van Andel, Submarine thermal springs on the Galapagos Rift, Science, 203, 1073-1083, 1979.

Craig, H., and R. Poreda, Papatua Expedition, Legs V and VI, SIO Ref. 87-14, 80 pp., 1987.

Craig, H., W. B. Clarke, and M. A. Beg, Excess 3He in deep water on the East Pacific Rise, Earth. Planet. Sci. Lett., 26, 125-132, 1975.

Craig, H., V. K. Craig, and K. -R. Kim, Papatua Expedition I: Hydrothermal vent surveys in Back-arc basins: the Lau, N. Fiji, Woodlark, and Manus Basins and Harve Trough, Eos Trans. AGU, 68, 100, 1987.

Crane, K., F. A. Aikman III, R. Embley, S. Hammond, A. Malahoff, and J. Lupton, The distribution of geothermal fields on the Juan de Fuca Ridge, J. Geophys. Res., 90, 727-744, 1985.

Crane, K., F. A. Aikman III, and J. P. Foucher, The distribution of geothermal fields along the East Pacific Rise from 13°10'N to 8°20'N: Implications for deep-seated origins, Mar. Geophys. Res., 9, 211-236, 1988.

DeMets, C., R. G. Gordon, D. F. Argus, and S. Stein, Current plate motions, Geophys. J. Int., 101, 425-478, 1990.

Detrick, R. S., P. Buhl, E. Vera, J. Mutter, J. Orcutt, J. Madsen, and T. Brocher, Multi-channel seismic imaging of a crustal magma chamber along the East Pacific Rise, Nature, 326, 35-41, 1987.

Detrick, R. S., A. J. Harding, G. M. Kent, J. A. Orcutt, J. C. Mutter, and P. Buhl, Seismic structure of the southern East Pacific Rise, Science, 259, 499-503, 1993.

Edmond, J. M., C. Measures, R. E. McDuff, L. H. Chan, R. Collier, B. Grant, L. I. Gordon, and J. B. Corliss, Ridge crest hydrothermal activity and the balances of major and minor elements in the ocean: The Galapagos data, Earth Planet. Sci. Lett., 46, 1-18, 1979.

Edmond, J. M., K. L. Von Damm, R. E. McDuff, and C. I. Measures, Chemistry of hot springs on the East Pacific Rise and their effluent dispersal, Nature, 297, 187-191, 1982.

Elderfield, H., C. German, M. Palmer, B. Murton, C. Chin, M. Greaves, E. Gurvich, R. James, G. Klinkhammer, E. Ludford, R. Mills, M. Rudnicki, J. Thomson, and A. Williams, Preliminary geochemical results from the Broken Spur hydrothermal field, 29°N, Mid-Atlantic Ridge, Eos Trans. AGU, 74(43 Supplement), 99, 1993.

Embley, R. W., W. W. Chadwick Jr., I. R. Jonasson, D. A. Butterfield, and E. T. Baker, Initial results of the rapid response to the 1993 CoAxial event: Relationships between hydrothermal and volcanic processes, Geophys. Res. Lett., 22(2), 43-46, 1995.

Feely, R. A., J. H. Trefry, G. J. Massoth, and S. Metz, A comparison of the scavenging of phosphorus and arsenic from seawater by hydrothermal iron oxyhydroxides in the Atlantic and Pacific Oceans, Deep-Sea Res., 38, 617-623, 1991.

Feely, R. A., J. F. Gendron, E. T. Baker, and G. T. Lebon, Hydrothermal plumes along the East Pacific Rise, 8°40' to 11°50'N: Particle distribution and composition, Earth Planet. Sci. Lett. 128, 19-36, 1994.

Feely, R. A., K. Marumo, T. Urabe, E. T. Baker, J. F. Gendron, and G. T. Lebon, Elemental chemistry of hydrothermal particles along the superfast-spreading southern East Pacific Rise, 13°50'-18°40'S, Eos Trans. AGU, 75(44 Supplement), 321, 1994.

Fouquet, Y., U. von Stackelberg, J. L. Charlou, J. P. Donval, J. P. Foucher, J. Erzinger, P. Herzig, R. Mühe, M. Wiedicke, S. Soakai, and H. Whitechurch, Hydrothermal activity in the Lau back-arc basin: Sulfides and water chemistry, Geology, 19, 303-306, 1991

Fox, C. G., Real-time detection of a volcanic eruption on the Juan de Fuca Ridge using the U.S. Navy's Sound Surveillance System, Geophys. Res. Lett., in press.

Gamo, T., H. Sakai, J. Ishibashi, E. Nakayama, K. Isshiki, H. Matsuura, K. Shitashima, K. Takeuchi, and S. Ohta, Hydrothermal plumes in the eastern Manus Basin, Bismark Sea: CH4, Mn, Al, and pH anomalies, Deep-Sea Res., 40, 2335-2349, 1993.

Gamo, T., E. Nakayama, H. Obata, K. Okamura, K. Isshiki, S. Kanayama, K. Shitashima, T. Oomori, T. Koizumi, and S. Matsumoto, Chemical evidence for the occurrence of hydrothermal activity on the Central Indian Ridge near the Rodriguez Triple Junction, Eos Trans. AGU, 75(44 Supplement), 314, 1994.

Gendron, J. F., J. P. Cowen, R. A. Feely, and E. T. Baker, Age estimate for the 1987 megaplume on the southern Juan de Fuca Ridge using excess radon and manganese partitioning, Deep-Sea Res., 40, 1559-1567, 1993.

German, C. R., A. C. Campbell, and J. M. Edmond, Hydrothermal scavenging at the Mid-Atlantic Ridge: Modification of trace element dissolved fluxes, Earth Planet. Sci. Lett., 107, 101-114, 1991.

German, C. R., J. Briem, C. Chin, M. Danielsen, S. Holland, R. James, A. Jónsdóttir, E. Ludford, C. Moser, J. Olafsson, M. R. Palmer, and M. D. Rudnicki, Hydrothermal activity on the Reykjanes Ridge: The Steinahóll Vent-field at 63°06'N, Earth Planet. Sci. Lett., 121, 647-654, 1994.

German, C. R., E. T. Baker, and G. Klinkhammer, The regional setting of hydrothermal activity, in Hydrothermal Vents and Processes, edited by D. Dixon, L. M. Parson, and C. L. Walker, Geol. Soc. London Spec. Pub. No. 87, 3-15, 1995.

German, C. R., B. A. Barreiro, N. C. Higgs, T. A. Nelsen, E. M. Ludford, and M. R. Palmer, Seawater-metasomatism in hydrothermal sediments, Escanaba Trough, Chem. Geol., in press.

Grimaud, D., J. I. Ishibashi, Y. Lagabrielle, J. M. Auzende, and T. Urabe, Chemistry of hydrothermal fluids from the 17°S active site on the North Fiji Basin Ridge (SW Pacific), Chem. Geol., 93, 209-218, 1991.

Hautala, S., and S. Riser, A nonconservative -spiral determination of the deep circulation in the eastern south Pacific, J. Phys. Oceanogr., 23, 1975-2000, 1993.

Haymon, R. M., D. J. Fornari, M. H. Edwards, S. Carbotte, D. Wright, and K. C. Macdonald, Hydrothermal vent distribution along the East Pacific Rise crest (9°09'-54'N) and its relationship to magmatic and tectonic processes on fast-spreading mid-ocean ridges, Earth Planet. Sci. Lett., 104, 513-534, 1991.

Haymon, R. M., D. J. Fornari, K. L. Von Damm, M. D. Lilley, M. R. Perfit, J. M. Edmond, W. C. Shanks, III, R. A. Lutz, J. M. Grebmeier, S. Carbotte, D. Wright, E. McLaughlin, M. Smith, N. Beedle, and E. Olson, Volcanic eruption of the mid-ocean ridge along the East Pacific Rise at 9°45'-52'N: Direct submersible observations of seafloor phenomena associated with an eruption event in April, 1991, Earth Planet. Sci. Lett., 119, 85-101, 1993.

Helfrich, K., and K. Speer, Oceanic hydrothermal circulation: mesoscale and basin-scale flow, in Physical, Chemical, Biological and Geological Interactions Within Hydrothermal Systems, this volume.

Herzig, P. M., and W. L. Plüger, Exploration for hydrothermal mineralization near the Rodriguez Triple Junction, Indian Ocean, Can. Mineral., 26, 721-736, 1988.

Horibe, Y., and H. Craig, Papatua Expedition III: Hydrothermal vents in the Mariana Trough and Kagoshima Bay (Sakurajima Volcano), Eos Trans. AGU, 68, 100, 1987.

Horibe, Y., K.-R. Kim, and H. Craig, Hydrothermal methane plumes in the Mariana back-arc spreading center, Nature, 324, 131-133, 1986.

Ishibashi, J.-I., T. Gamo, H. Sakai, Y. Nojiri, G. Igarashi, K. Shitashima, and H. Tsubota, Geochemical evidence for hydrothermal activity in the Okinawa Trough, Geochem. J., 22, 107-114, 1988.

Ishibashi, J., H. Wakita, K. Okamura, E. Nakayama, R. A. Feely, G. J. Massoth, and E. T. Baker, Methane and manganese distribution in hydrothermal plumes along the East Pacific Rise, 13°50'-18°40'S, Eos Trans. AGU, 75(44 Supplement), 321, 1994.

James, R. H., H. Elderfield, M. D. Rudnicki, C. R. German, M. R. Palmer, C. Chin, M. J. Greaves, E. Gurvich, G. P. Klinkhammer, E. Ludford, R. A. Mills, J. Thomson, and A. C. Williams, Hydrothermal plumes at Broken Spur, 29°N, Mid-Atlantic Ridge: chemical and physical characteristics, in Hydrothermal Vents and Processes, edited by D. Dixon, L. M. Parson, and C. L. Walker, Geol. Soc. London Spec. Pub., in press a.

James, R. H., H. Elderfield, and M. R. Palmer, The chemistry of vent fluids from the Broken Spur hydrothermal site, 29°N Mid-Atlantic Ridge, Geochim. Cosmochim. Acta, in press, b.

Jamous, D., L. Mémery, C. Andríe, P. Jean-Baptiste, and L. Merlivat, The distribution of helium 3 in the deep western and southern Indian Ocean, J. Geophys. Res., 97, 2243-2250, 1992.

Johnson, K. S., C. L. Beehler, and C. M. Sakamoto-Arnold, A submersible flow analysis system, Anal. Chim. Acta., 179, 245-257, 1986a.

Johnson, K. S., C. L. Beehler, C. M. Sakamoto-Arnold, and J. Childress, In situ measurements of chemical distributions in a deep-sea hydrothermal vent field, Science, 231, 1139-1141, 1986b.

Jones, C. J., H. P. Johnson, and J. R. Delaney, Distribution of hydrothermal manganese over the Juan de Fuca Ridge, Geophys. Res. Lett., 8, 873-876, 1981.

Kadko, D., An assessment of the effect of chemical scavenging within submarine hydrothermal plumes upon ocean geochemistry, Earth. Planet. Sci. Lett., 361-374, 1994.

Kadko, D., J. Baross, and J. Alt, The magnitude and global impact of hydrothermal flux, in Physical, Chemical, Biological, and Geological Interactions within Hydrothermal Systems, this volume.

Klinkhammer, G., Fiber optic spectrometers for in situ measurement in the oceans: the ZAPS probe, Mar. Chem., 47, 13-20, 1994.

Klinkhammer, G., and A. Hudson, Dispersal patterns for hydrothermal plumes in the South Pacific using manganese as a tracer, Earth Planet. Sci. Lett., 79, 241-249, 1986.

Klinkhammer, G., M. Bender, and R. F. Weiss, Hydrothermal manganese in the Galapagos Rift, Nature, 269, 319-320, 1977.

Klinkhammer, G., H. Elderfield, and A. Hudson, Rare earth elements in seawater near hydrothermal vents, Nature, 305, 185-188, 1983.

Klinkhammer, G., P. Rona, M. Greaves, and H. Elderfield, Hydrothermal manganese plumes in the mid-Atlantic Ridge rift valley, Nature, 314, 727-731, 1985.

Klinkhammer, G., H. Elderfield, M. Greaves, P. Rona, and T. Nelsen, Manganese geochemistry near high-temperature vents in the Mid-Atlantic rift valley, Earth Planet. Sci. Lett., 80, 230-240, 1986.

Klinkhammer, G. P., C. S. Chin, M. Rudnicki, C. Wilson, and C. R. German, Zapping coastal waters for dissolved manganese and blue fluorescence: Indicators of runoff, mixing, and exchange, Eos Trans. AGU, 75(44 Supplement), 327, 1994.

Klinkhammer, G., C. S. Chin, C. Wilson, and C. R. German, Venting from the Mid-Atlantic Ridge at 37°17'N: The Lucky Strike hydrothermal site, in Hydrothermal Vents and Processes, edited by D. Dixon, L. M. Parson, and C. L. Walker, Geol. Soc. London Spec. Pub., in press.

Knauss, J. A., On some aspects of the deep circulation of the Pacific, J. Geophys. Res., 67, 3943-3954, 1962.

Krasnov, S. G., I. I. Kreyter, and I. M. Poroshina, The distribution of hydrothermal vents on the East Pacific Rise (21°20'-22°40'S) based on a study of dispersion patterns of hydrothermal plumes, Oceanology, 32, 375-381, 1992.

Langmuir, C. H., D. Fornari, D. Colodner, J. L. Charlou, I. Costa, D. Desbruyeres, D. Desonie, T. Emerson, A. Fiala-Medoni, Y. Fouquet, S. Humphris, L. Saldanha, R. Sours-Page, M. Thatcher, M. Tivey, C. Van Dover, K. Von Damm, K. Wiese, and C. Wilson, Geological setting and characteristics of the Lucky Strike vent field at 37°17'N on the Mid-Atlantic Ridge, Eos Trans. AGU, 74(43 Supplement), 99, 1993.

Lavelle, J. W., A convection model for hydrothermal plumes in a cross flow, NOAA Tech. Memo. ERL PMEL-102, 18 pp., 1994.

Lilley, M. D., R. A. Feely, and J. H. Trefry, Chemical and biochemical transformations in hydrothermal plumes, in Physical, Chemical, Biological, and Geological Interactions Within Hydrothermal Systems, this volume.

Little, S. A., K. D. Stolzenbach, and R. P. Von Herzen, Measurements of plume flow from a hydrothermal vent, J. Geophys. Res., 92, 2587-2596, 1987.

Lisitsyn, A. P., K. A. W. Crook, Yu. A. Bogdanov, L. P. Zonenshayn, K. G. Murav'yev, W. Tufar, Ye. G. Gurvich, V. V. Gordeyev, and G. V. Ivanov, A hydrothermal field in the rift zone of the Manus Basin, Bismark Sea, Internat. Geol. Rev., 35, 105-126, 1993.

Lonsdale, P., Abyssal circulation of the southeastern Pacific and some geological implications, J. Geophys. Res., 81, 1163-1176, 1976.

Lupton, J. E., Helium-3 in the Guaymas Basin: Evidence for injection of mantle volatiles in the Gulf of California, J. Geophys. Res., 84, 7446-7452, 1979.

Lupton, J. E., Water column hydrothermal plumes on the Juan de Fuca Ridge, J. Geophys. Res., 95, 12,829-12,842, 1990.

Lupton, J. E., Hydrothermal plumes: near and far field, in Physical, Chemical, Biological, and Geological Interactions Within Hydrothermal Systems, this volume.

Lupton, J. E., and H. Craig, A major helium-3 source at 15°S on the East Pacific Rise, Science, 214, 13-18, 1981.

Lupton, J. E., J. R. Delaney, H. P. Johnson, and M. K. Tivey, Entrainment and vertical transport of deep-ocean water by buoyant hydrothermal plumes, Nature, 316, 621-623, 1985.

Lupton, J. E., E. T. Baker, M. J. Mottl, F. J. Sansone, C. G. Wheat, J. A. Resing, G. J. Massoth, C. I. Measures, and R. A. Feely, Chemical and physical diversity of hydrothermal plumes along the East Pacific Rise, 8°45'N to 11°50'N, Geophys. Res. Lett., 20, 2913-2916, 1993.

Lupton, J. E., R. Greene, G. Paradis, T. Urabe, and E. Baker, The far-field hydrothermal plume from the southern East Pacific Rise and its relationship to sources on the ridge axis, Eos Trans. AGU, 75(44 Supplement), 320, 1994.

Massoth, G. J., H. B. Milburn, K. S. Johnson, K. H. Coale, M. F. Stapp, C. Meinig, and E. T. Baker, A SUAVE (Submersible System Used to Assess Vented Emissions) approach to plume sensing: the Buoyant Plume Experiment at Cleft Segment, Juan de Fuca Ridge and plume exploration along the EPR 9-11°N, Eos Trans. AGU, 72(44 Supplement), 234, 1991.

Massoth, G. J., E. T. Baker, R. A. Feely, K. Okamura, and E. Nakayama, Hydrothermal sources and fluxes along the southern East Pacific Rise, 13°50'-18°40'S: Thermochemical insights from plume data, Eos Trans. AGU, 75(44 Supplement), 321, 1994.

McConachy, T. F., and S. D. Scott, Real-time mapping of hydrothermal plumes over southern Explorer Ridge, NE Pacific Ocean, Mar. Mining, 6, 181-204, 1987.

McDougall, T. J., Bulk properties of "hot smoker" plumes, Earth Planet. Sci. Lett., 99, 185-194, 1990.

McDuff, R., Physical dynamics of deep-sea hydrothermal plumes, in Physical, Chemical, Biological, and Geological Interactions Within Hydrothermal Systems, this volume.

Metz, S., and J. H. Trefry, Field and laboratory studies of metal uptake and release by hydrothermal precipitates, J. Geophys. Res., 98, 9661-9666, 1993.

Michard, G., F. Albarède, A. Michard, J. F. Minster, J. L. Charlou, and N. Tan, Chemistry of solutions from the 13°N East Pacific Rise hydrothermal site, Earth Planet. Sci. Lett., 67, 297-307, 1984.

Middleton, J. M., and R. E. Thomson, Modeling the rise of hydrothermal plumes, Can. Tech. Rpt. Hydro. and Ocean Sci., No. 69, Canadian Dept. of Fisheries and Oceans, Sidney, B.C., 18 pp., 1986.

Mottl, M. J., and T. F. McConachy, Chemical processes in buoyant hydrothermal plumes on the East Pacific Rise near 21°N, Geochim. Cosmochim. Acta, 54, 1911-1927, 1990.

Mottl, M. J., F. T. Sansone, C. G. Wheat, J. A. Resing, E. T. Baker, and J. E. Lupton, Hydrothermal plumes along the East Pacific Rise, 8°40' to 11°50'N: Dissolved manganese and methane, Geochim. Cosmochim. Acta, 59(20), 4147-4165, 1995.

Morton, B. R., G. I. Taylor, and J. S. Turner, Turbulent gravitational convection from maintained and instantaneous sources, Proc. Royal Soc. London, A234, 1-23, 1956.

Mullineaux, L., P. B. Weibe, and E. T. Baker, Larvae of benthic invertebrates in hydrothermal vent plumes over Juan de Fuca Ridge, Mar. Biol., 122, 585-596, 1995.

Murton, B. J., G. Klinkhammer, K. Becker, A. Briais, D. Edge, N. Hayward, N. Millard, I. Mitchell, I. Rouse, M. Rudnicki, K. Sayanagi, H. Sloan, and L. Parson, Direct evidence for the distribution and occurrence of hydrothermal activity between 27°N-30°N on the Mid-Atlantic Ridge, Earth Planet. Sci. Lett., 125, 119-128, 1994.

Nelsen, T. A., G. P. Klinkhammer, J. H. Trefry, and R. P. Trocine, Real-time observation of dispersed hydrothermal plumes using nephelometry: examples from the Mid-Atlantic Ridge, Earth. Planet. Sci. Lett., 81, 245-252, 1986/87.

Nojiri, Y., J. Ishibashi, T. Kawai, and H. Sakai, Hydrothermal plumes along the North Fiji Basin spreading axis, Nature, 342, 667-670, 1989.

Ocean Drilling Program Leg 106 Scientific Party, Drilling the Snake Pit hydrothermal sulfide deposit on the Mid-Atlantic Ridge, lat 23°22'N, Geology, 14, 1004-1007, 1986.

Phipps Morgan, J., and Y. J. Chen, The genesis of oceanic crust: Magma injection, hydrothermal circulation, and crustal flow, J. Geophys. Res. 98, 6283-6297, 1993.

Plüger, W. L., P. M. Herzig, K. P. Becker, G. Deissmann, D. Schöps, J. Lange, A. Jenisch, S. Ladage, H. H. Richnow, T. Schulze, and W. Michaelis, Discovery of hydrothermal fields at the Central Indian Ridge, Mar. Mining, 9, 73-86, 1990.

Purdy, G. M., L. S. L. Kong, G. L. Christeson, and S. C. Solomon, Relationship between spreading rate and seismic structure of mid-ocean ridges, Nature, 355, 815-817, 1992.

Reid, J. L., Evidence of an effect of heat flux from the East Pacific Rise upon the characteristics of the mid-depth waters, Geophys. Res. Lett., 9, 381-384, 1982.

RISE Project Group, East Pacific Rise: Hot springs and geophysical experiments, Science, 207, 1421-1433, 1980.

Rona, P. A., G. Klinkhammer, T. A. Nelsen, J. H. Trefry, and H. Elderfield, Black smokers, massive sulfides and vent biota at the mid-Atlantic Ridge, Nature, 321, 33-37, 1986.

Rosenberg, N. D., J. E. Lupton, D. Kadko, R. Collier, M. D. Lilley and H. Pak, Estimation of heat and chemical fluxes from a seafloor hydrothermal vent field using radon measurements, Nature, 334, 604-607, 1988.

Rudnicki, M. D., and H. Elderfield, Theory applied to the Mid-Atlantic Ridge hydrothermal plumes: the finite difference approach, J. Volcanol. Geothermal Res., 50, 161-172, 1992.

Rudnicki, M. D., and H. Elderfield, A chemical model of the buoyant and neutrally buoyant plume above the TAG vent field, 26°N, Mid-Atlantic Ridge, Geochim. Cosmochim. Acta, 57, 2939-2957, 1993.

Sakai, H., T. Gamo, E.-S. Kim, K. Shitashima, F. Yanagisawa, M. Tsutsumi, J. Ishibashi, Y. Sano, H. Wakita, T. Tanaka, T. Matsumoso, T. Naganuma, and K. Mitsuzawa, Unique chemistry of the hydrothermal solution in the mid-Okinawa Trough backarc basin, Geophys. Res. Lett., 17, 2133-2136, 1990.

Sedwick, P. N., T. Gamo, and G. M. McMurtry, Manganese and methane anomalies in the North Fiji Basin, Deep-Sea Res., 37, 891-896, 1990.

Sinton, J. M., S. M. Smaglik, J. J. Mahoney, and K. C. Macdonald, Magmatic processes at superfast mid-ocean ridges: Glass compositional variations along the East Pacific Rise, 13°-23°S, J. Geophys. Res., 96, 6133-6155, 1991.

Speer, K. G., and P. A. Rona, A model of an Atlantic and Pacific hydrothermal plume, J. Geophys. Res., 94, 6213-6220, 1989.

Taylor, B., Bismark Sea: Evolution of a back-arc basin, Geology, 7, 171-174, 1979.

Taylor, B., and G. D. Karner, On the evolution of marginal basins, Rev. Geophys. Space Phys., 21, 1727-1741, 1983.

Thomson, R. E., J. R. Delaney, R. E. McDuff, D. R. Janecky, and J. S. McClain, Physical characteristics of the Endeavour Ridge hydrothermal plume during July 1988, Earth Planet. Sci. Lett., 111, 141-154, 1992.

Tunnicliffe, V., M. Botros, M. E. de Burgh, A. Dinet, H. P. Johnson, S. K. Juniper, and R. E. McDuff, Hydrothermal vents of Explorer Ridge, northeast Pacific, Deep-Sea Res., 33, 401-412, 1986.

Turner, J. S., Buoyancy Effects in Fluids, 367 pp., Cambridge Univ. Press, New York, 1973.

Urabe, T., E. T. Baker, and Ridge Flux Group, An overview of multidisciplinary cruise of R/V Melville to superfast-spreading East Pacific Rise, 13.5°S-18.5°S, Eos Trans. AGU, 75(44 Supplement), 320, 1994.

Von Damm, K. L., Seafloor hydrothermal activity: Black smoker chemistry and chimneys, Ann. Rev. Earth Planet. Sci., 18, 173-204, 1990.

Von Damm, K. L., Controls on the chemistry and temporal variability of seafloor hydrothermal fluids, in Physical, Chemical, Biological, and Geological Interactions Within Hydrothermal Systems, this volume.

Von Damm, K. L., and J. L. Bischoff, Chemistry of hydrothermal solutions from the southern Juan de Fuca Ridge, J. Geophys. Res., 92, 11,334-11,346, 1987.

Von Damm, K. L., J. M. Edmond, C. I. Measures, and B. Grant, Chemistry of submarine hydrothermal solutions at Guaymas Basin, Gulf of California, Geochim. Cosmochim. Acta, 49, 2221-2237, 1985a.

Von Damm, K. L., J. M. Edmond, B. Grant, C. I. Measures, B. Walden, and R. F. Weiss, Chemistry of submarine hydrothermal solutions at 21°N, East Pacific Rise, Geochim. Cosmochim. Acta, 49, 2197-2220, 1985b.

Von Damm, K. L., J. M. Grebmeier, and J. M. Edmond, Preliminary chemistry of hydrothermal vent fluids from 9-10°N East Pacific Rise, Eos Trans. AGU, 72(44 Supplement), 480, 1991.

Von Herzen, R. P., Heat-flow values from the southeastern Pacific, Nature, 183, 882-883, 1959.

Warren, B. A., Transpacific hydrographic sections at Lats. 43°S and 28°S: the SCORPIO Expedition--II, Deep-Sea Res., 20, 9-38, 1973.

Welhan, J. A., and H. Craig, Methane, hydrogen and helium in hydrothermal fluids at 21°N on the East Pacific Rise, in Hydrothermal processes at seafloor spreading centers, edited by P. A. Rona, K. Boström, L. Laubier, and K. L. Smith, Jr., Plenum Press, pp. 391-410, 1983.

Wilson, C., K. Speer, J. L. Charlou, H. Bougault, and G. Klinkhammer, Mid-Atlantic Ridge hydrography and anomalies at Lucky Strike, J. Geophys. Res., in press.


Return to Abstract

PMEL Outstanding Papers

PMEL Publications Search

PMEL Homepage