undersea exploration

undersea exploration

Introduction

      the investigation and description of the ocean waters and the seafloor and of the Earth beneath.

Primary objectives and accomplishments
      Included in the scope of undersea exploration are the physical and chemical properties of seawater, all manner of life in the sea, and the geological and geophysical features of the Earth's crust. Researchers in the field define and measure (measurement) such properties; prepare maps in order to identify patterns; and utilize these maps, measurements, and theoretical models to achieve a better grasp of how the Earth works as a whole. This knowledge enables scientists to predict, for example, long-term weather and climatic changes and leads to more efficient exploration and exploitation of the Earth's resources, which in turn result in better management of the environment in general.

      The multidisciplinary expedition (Challenger Expedition) of the British ship “Challenger” in 1872–76 was the first major undersea survey. Although its main goal was to search for deep-sea life by means of net tows and dredging, the findings of its physical and chemical studies expanded scientific knowledge of temperature and salinity distribution of the open seas. Moreover, depth measurements by wire soundings were carried out all over the globe during the expedition.

      Since the time of the “Challenger” voyage, scientists have learned much about the mechanics of the ocean, what it contains, and what lies below its surface. Investigators have produced global maps showing the distribution of surface winds as well as of heat and rainfall, which all work together to drive the ocean in its unceasing motion. They have discovered that storms at the surface can penetrate deep into the ocean and, in fact, cause deep-sea sediments to be rippled and moved. Recent studies also have revealed that storms called eddies (eddy) occur within the ocean itself and that such a climatic anomaly as El Niño is caused by an interaction of the ocean and the atmosphere.

      Other investigations have shown that the ocean absorbs large amounts of carbon dioxide and hence plays a major role in delaying its buildup in the atmosphere. Without the moderating effect of the ocean, the steadily increasing input of carbon dioxide into the atmosphere (due to the extensive burning of coal, oil, and natural gas) would result in the rapid onset of the so-called greenhouse effecti.e., a warming of the Earth caused by the absorption and reradiating of infrared energy to the terrestrial surface by carbon dioxide and water vapour in the air.

      The field of marine biology has benefitted from the development of new sampling methods. Among these, broad ranging acoustical techniques have revealed diverse fish populations and their distribution, while direct, close up observation made possible by deep-sea submersibles has resulted in the discovery of unusual (and unexpected) species and phenomena.

      In the area of geology (marine geology), undersea exploration of the topography of the seafloor and its gravitational and magnetic properties has led to the recognition of global patterns of continental plate motion. These patterns form the basis of the concept of plate tectonics, which synthesized earlier hypotheses of continental drift and seafloor spreading. As noted earlier, this concept not only revolutionized scientific understanding of the Earth's dynamic features (e.g., seismic activity, mountain-building, and volcanism) but also yielded discoveries of economic and political impact. Earth scientists found that the mid-ocean centres of seafloor spreading also are sites of important metal deposits. The hydrothermal circulations associated with these centres produce sizable accumulations of metals important to the world economy, including zinc, copper, lead, silver, and gold. Rich deposits of manganese, cobalt, nickel, and other commercially valuable metals have been found in nodules distributed over the entire ocean floor. The latter discovery proved to be a major factor in the establishment of the Convention of the Law of the Sea (1982), which calls for the sharing of these resources among developed and developing nations alike. Exploitation of these findings awaits only the introduction of commercially viable techniques for deep-sea mining and transportation.

Basic elements of undersea exploration

Platforms
      Undersea exploration of any kind must be conducted from platforms, in most cases, ships, buoys, aircraft, or satellites. Typical oceanographic vessels capable of carrying out a full complement of underwater exploratory activities range in size from about 50 to 150 metres. They support scientific crews of 16 to 50 persons and generally permit a full spectrum of interdisciplinary studies. One example of a research vessel of this kind is the “Melville,” operated by the Scripps Institution of Oceanography. It has a displacement of 2,075 tons and can carry 25 scientists in addition to 25 crew members. It is powered by a dual cycloidal propulsion system, which provides remarkable manoeuvrability.

 The “JOIDES Resolution,” operated by Texas A & M University for the Joint Oceanographic Institutions for Deep Earth Sampling, represents a major advance in research vessels. A converted commercial drill ship, it measures 145 metres in length, has a displacement of 18,600 tons, and is equipped with a derrick that extends 62 metres above the waterline (see photograph—>). A computer-controlled dynamic positioning system enables the ship to remain over a specific location while drilling in water to depths as great as 8,300 metres. The drilling system of the ship is designed to collect cores from below the ocean floor; it can handle 9,200 metres of drill pipe. The vessel thus can sample most of the ocean floor, including the bottoms of deep ocean basins and trenches. The “JOIDES Resolution” has other notable capabilities. It can operate in waves as high as eight metres, winds up to 23 metres per second, and currents as strong as 1.3 metres per second. It has been outfitted for use in ice so that it can conduct drilling operations in high latitudes. The ship can accommodate 50 scientists as well as the crew and drilling team, and its geophysical laboratories total nearly 930 square metres.

      Other specialized vessels include the deep submergence research vehicle known as “Alvin,” which can carry a pilot and two scientific observers to a depth of 4,000 metres. The manoeuvrability of the “Alvin” was pivotal to the discoveries of the mineral deposits at the mid-ocean seafloor spreading centres and of previously unknown biological communities living at those sites. Another versatile vessel is the Floating Instrument Platform (FLIP). It is a long narrow platform that is towed in a horizontal position to a research site. Once on location, the ballast tanks are flooded to flip the ship to a vertical position. Only 17 metres of the ship extend above the waterline, with the remaining 92 metres completely submerged. The rise and fall of the waves cause a very small change in the displacement, resulting in a high degree of stability.

      New ship designs that promise even greater stability and ease of use include that of the Small Waterplane Area Twin Hull (SWATH) variety. This design type requires the use of twin submerged, streamlined hulls to support a structure that rides above the water surface. The deck shape is entirely unconstrained by the hull shape, as is the case for conventional surface vessels. Ship motion is greatly reduced because of the depth of the submerged hulls. For a given displacement, a SWATH-type vessel can provide twice the amount of deck space that a single-hull ship can, with only 10 percent of the motion of the single-hull design type. In addition, a large centre opening, or well, can be used to display and recover instruments.

      Exploration of any kind is useful only when the location of the discoveries can be noted precisely. Thus, navigation has always been a key to undersea exploration.

      There are various ways by which the position of a vessel at sea can be determined. In cases where external references such as stars or radio and satellite beacons are unavailable or undetectable, inertial navigation (inertial guidance system), which relies on a stable gyroscope for determining position, is commonly employed. It is far more accurate than the long-used technique of dead reckoning, which is dependent on a knowledge of the ship's original position and the effects of the winds and ocean currents on the vessel.

      Another modern position-fixing method is all-weather, long-range radio navigation. It was introduced during World War II as loran (long-range navigation) A, a system that determines position by measuring the difference in the time of reception of synchronized pulses from widely spaced transmitting stations. The latest version of this system, Loran C, uses low-frequency transmissions and derives its high degree of accuracy from precise time-difference measurements of the pulsed signals and the inherent stability of signal propagation. Users of Loran C are able to identify a position with an accuracy of 0.4 kilometre and a repeatability of 15 metres at a distance of up to about 2,220 kilometres from the reference stations. The Loran C system covers heavily travelled regions in the North Pacific and North Atlantic oceans, parts of the Indian Ocean, and the Mediterranean Sea.

      Satellite navigation has proved to be the most accurate method of locating geographical position. A polar-orbiting satellite system called Transit was established in the early 1960s by the United States to provide global coverage for ships at sea. In this system, a vessel pinpoints its position relative to a set of satellites whose orbits are known by measuring the Doppler shift (Doppler effect) of a received signal—i.e., the change in the frequency of the received signal from that of the transmitted signal. The Transit system suffers from one major drawback. Because of the limited number of system satellites, the frequency with which position determinations can be made each day is relatively low, particularly in the tropics. The system is being improved to provide nearly continuous positioning capability at sea. This expanded version, the Global Positioning System (GPS) (GPS), is to have 18 satellites, six in each of three orbital planes spaced 120° apart. The GPS is designed to provide fixes anywhere on Earth to an accuracy of 20 metres and a relative accuracy 10 times greater.

Methodology and instrumentation

Water sampling for temperature and salinity
      The temperature, chemical environment, and movement and mixing of seawater are fundamental to understanding the physical, chemical, and biological features of the ocean and the geology of the ocean floor. Traditionally, oceanographers have collected seawater by means of specially adapted water-sampling bottles. The most universal water sampler used today, the Nansen bottle, is a modification of a type developed in the latter part of the 19th century by the Norwegian Arctic explorer and oceanographer Fridtjof Nansen (Nansen, Fridtjof). It is a metal sampler equipped with special closing valves that are actuated when the bottle, attached by one end to a wire that carries it to the desired depth, rotates about that end. A mercury thermometer fastened to the bottle records the temperature at the specified depth. The design of the device is such that, when it is inverted, its mercury column breaks. The amount of mercury remaining in the graduated capillary portion of the thermometer indicates the temperature at the point of inversion. This type of reversing thermometer and the Nansen bottle are extensively used by oceanographers because of their accuracy and dependability in a harsh environment.

      The temperature and salinity of the ocean have been mapped with data gathered by many ships over many years. This information is used for tracing heat and water movement and mixing, as well as for making density measurements, which are employed in calculating ocean currents. It was noted as early as the “Challenger” expedition that the salt dissolved in seawater has remarkably constant major constituents. As a consequence, it is possible to map water density patterns within the sea with measurements of only the water temperature and one major property of the sea salt (e.g., the chloride ion content or the electrical conductivity) to arrive at an accurate estimate of the density of a given sample.

      Standard laboratory techniques such as titration are routinely used at sea for determining chlorinity. Chlorinity can be briefly defined as the number of grams of chlorine, bromine, and iodine contained in one kilogram of seawater, assuming that the bromine and iodine are replaced by chlorine. Salinity is the total weight of dissolved solids, in grams, found in one kilogram of seawater and may be determined from the concentration of chlorinity because of the constancy of major constituents. In the traditional technique, a solution of silver nitrate of a known strength is added to a sample of seawater to produce the same reaction as with “standard” seawater. The difference in the amounts added gives the degree of chlorinity. To ensure worldwide uniformity in chlorinity and salinity determinations, the International Council for the Exploration of the Sea prepared a universal reference, Eau de Mer Normale (“Standard Seawater”), in 1902. A new primary standard, prepared in 1937 and having a chlorinity of 19.381 parts per 1,000, is used to determine the chlorinities of all batches of standard seawater. It also is utilized to calibrate electrical conductivity measurements (see below).

      Accurate and continuous measurements of temperature as it changes with depth are required for understanding how the ocean moves and mixes heat. To provide the necessary detail, temperature profilers had to be developed; then, with the introduction of reliable conductivity sensors, salinity profilers were added. An instrument called the bathythermograph (BT), which has been used since the early 1940s to obtain a graphic record of water temperature at various depths, can be lowered from a ship while it is moving at reduced speed. In this instrument a depth element (pressure-operated bellows) drives a slide of smoked glass or metal at right angles to a stylus. Actuated by a thermal element (liquid-filled bourdon tube) that expands and contracts in response to changes in temperature, the stylus scribes a continuous record of temperature and depth.

      An expendable bathythermograph (XBT) was developed during the 1970s and has come into increasingly wider use. Unlike the BT, this instrument requires an electrical system aboard the research platform. It detects temperature variations by means of a thermistor (an electrical resistance element made of a semiconductor material) and depends on a known fall rate for depth determination. The sensor unit of the XBT is connected to the research platform by a leak-proof, insulated two-conductor cable. This cable is wound around a pair of large spools in an arrangement resembling that of a fisherman's spinning reel. In operation, the cable is unwound from each of the spools in a direction that is parallel to the axis of the respective spool. As a result, the cable unwinds from both the platform—either a ship or an airplane—and the sensor unit simultaneously but independently. Because of this double-spool arrangement, the sensor unit can free-fall from wherever it hits the sea surface and is completely unaffected by the direction or speed of the craft from which it was deployed. One of the principal reasons why the XBT has proved so useful is that it can provide a record of considerable depth even when it is deployed from a ship moving at full speed.

      Until the late 1950s, salinity was universally determined by titration. Since then, shipboard electrical conductivity systems have become widely used. Salinity-Temperature-Depth (STD) and the more recent Conductivity-Temperature-Depth (CTD) systems have greatly improved on-site hydrographic sampling methods. They have enabled oceanographers to learn much about small-scale temperature and salinity distributions.

      The most recent version of the CTD systems features rapid-response conductivity and temperature sensors. The conductivity sensor consists of a tiny cell with four platinum electrodes. This type of conductivity cell virtually eliminates errors resulting from the polarization that occurs where the electrodes come in contact with seawater. The temperature sensor combines a tiny thermistor with a platinum-resistance thermometer. Its operations are carried out in such a way as to fully exploit the fast response of the thermistor and the high accuracy of the platinum thermometer. In addition, the system uses a strain gauge as a pressure sensor, the gauge being adjusted to reduce temperature effects to a minimum. This CTD system is extremely reliable. While its temperature precision is greater than 0.001° C over a range of −3° to +32° C, its conductivity precision is on the order of one part per million.

      Electrical conductivity measurement of seawater salinity has been so effective that it has given rise to a new practical salinity scale, one that is defined on the basis of conductivity ratio. This scale has proved to be a more reliable way of determining density (i.e., the weight of any given volume of seawater at a specified temperature) than the chlorinity scale traditionally used. Such is the case because chlorinity is ion specific while conductivity is sensitive to changes in any ion. Investigators have found that measurements of conductivity ratio make it possible to predict density with a precision almost one order of magnitude greater than was permitted by the chlorinity measurements of the past.

Water sampling for chemical constituents
      Nutrient concentration (e.g., phosphate, nitrate, silicate), the pH (acidity), and the proportion of dissolved gases are used by the ocean chemist to determine the age, origin, and movement of water masses and their effect on marine life. Analysis of dissolved gases, for example, is useful in tracing ocean mixing, in studying gas production in the ocean, and in elucidating the natural cycles of atmospheric pollutants. Many such measurements are conducted aboard ship by autoanalyzers, devices that continually monitor a flow of seawater by spectral techniques. Those analyses that cannot be accomplished by an autoanalyzer are carried out with discrete samples in shipboard or shore-based laboratories.

      Radioactive chemical tracers are of special interest. Radioisotopes (radioactive isotope) serve as time clocks, thus offering a means of determining the age of water masses, the absolute rates of oceanic mixing, and the generation and destruction of plant tissue. The distribution of these time clocks is controlled by the interaction of physical and biological processes, and so these influences must be disentangled before the clocks can be read. A notable example is the use of carbon-14 (14C). Today, a number of oceanographic laboratories make carbon-14 measurements of oceanic dissolved carbon for the study of mixing and transport processes in the deep ocean. Until recently large samples of water—about 200 litres (one litre = 0.264 gallon)—were required for analysis. New techniques use a linear accelerator (a device that greatly increases the velocity of electrically charged atomic and subatomic particles) as a sophisticated mass spectrometer to directly determine abundancy ratios of carbon-14/carbon-13/carbon-12 atoms. The advantage of the newer methodology is that only very small sample amounts—about 250 millilitres (one millilitre = 0.034 fluid ounce), are required for high accuracy measurements.

Measurements of ocean currents (ocean current)
      Ocean currents can be measured indirectly through data on density and directly with current meters. In the indirect technique, water density is computed from temperature and salinity observations, and pressure is then calculated from density. The resulting highs and lows of ocean pressure can be used to estimate ocean currents. The indirect technique establishes currents relative to a particular pressure surface; it is best for large-scale, low-frequency currents.

      Direct measurement of currents is used to establish absolute currents and to monitor rapidly varying changes. In order to measure currents directly, a current meter must accurately record the speed and direction of flow, and the platform or mooring has to be reliable, readily deployable, and extremely sturdy. Researchers are able to make continuous measurements of currents at levels below the surface layer for periods of more than a year.

      A typical system for the direct measure of ocean currents has three principal components: a surface or near-surface float; a line consisting of segments of wire and nylon that holds the current meters; and a release mechanism, signalled acoustically, which will drop an anchor when the system is ready to be brought back. A current meter typically employs a rotor equipped with a small direction vane that moves freely in line with the meter.

      One of the most important advances in modern instrument design has been the introduction of low-power, solid-state microelectronics. The accuracy of the Vector Averaging Current Meter (VACM), for example, has been improved appreciably by the use of integrated circuits, as has its data-handling capability. Because of the latter, the VACM can sample the direction and speed of currents roughly eight times during each revolution of the rotor. It then computes the north and east components of speed and stores this data, together with direction and time measurements, on a compact cassette recorder. The VACM is capable of making accurate measurements in wave fields as well as from moorings at the ocean surface because of its direct vector-averaging feature.

      Currents also can be measured by drifting floats, either at the surface or at a given depth. Tracking the location of the floats is critical. Surface floats can be followed by satellite, but subsurface drifters must be tracked acoustically. A drifter of this sort acts as an acoustical source and transmits signals that can be followed by a ship with hydrophones suspended into the sea. For such tracking, a low sound frequency is crucial because the higher the frequency of sound, the more rapidly is its energy absorbed by the sea. The longest range floats available during the mid-1980s operated at a frequency as low as about 250 hertz. Long-range floats usually drift along channels known as sound fixing and ranging (SOFAR) channels, which occur in various areas of the ocean where a particular combination of temperature and pressure conditions affect the speed of sound. In a sense, the SOFAR channel acts as a type of acoustic waveguide that focusses sound; as a consequence, several watts of sound can be detected as far away as 2,000 kilometres or so.

      Measuring vertical velocity in the ocean posed a major problem for years because of the difficulty of devising a platform that does not move vertically. During the 1960s oceanographers finally came up with a solution: they employed a neutrally buoyant float for measuring vertical velocities. This form of vertical-current meter consists of a cylindrical float on which fins are mounted at an angle. When water moves past the float, it causes the float to turn on its axis. Measurement of the rotation in relation to a compass yields the amount of vertical water movement.

      An extension of the neutrally buoyant float is the self-propelled, guided float. One such system, called a Self-Propelled Underwater Research Vehicle (SPURV), manoeuvres below the surface of the sea in response to acoustic signals from the research vessel. It can be used to produce horizontal as well as vertical profiles of various physical properties.

      A Doppler-sonar system for measuring upper-ocean current velocity transmits a narrow beam that scatters off drifting plankton and other organisms in the uppermost strata of the ocean. From the Doppler shift of the backscattered sound, the component of water velocity parallel to the beam can be determined to a range of 1,400 metres from the transmitter with a precision of one centimetre per 0.1 second (one centimetre = 0.394 inch).

      Integral to a complete picture of the ocean is a profile of velocity. Various methods have been devised for measuring currents as dependent on and varying with depth or horizontal position. Three techniques have been developed to make such measurements. The first involves acoustically tracking a “sinking float” as it descends toward the seafloor. The second technique entails the use of a free-fall device equipped with a current sensor. The third involves a class of current meter specially designed to move up and down a fixed line attached to a vessel, mooring, or drifting buoy. One such instrument has a roller block that couples the front of the instrument to a wire from the vessel. In this way, the motion of the vessel is decoupled from that of the instrument. Another important component of this instrument is its hull, a structure that not only furnishes buoyancy but also serves as a direction vane. In the bottom of the hull is a device that records velocity, temperature, and depth. The entire system descends at a rate of approximately 10 centimetres per second, resulting in a vertical resolution of several metres for the velocity profile produced.

Acoustic and satellite sensing
      Remote sensing of the ocean can be done by aircraft and Earth-orbiting satellites or by sending acoustic signals through it. These techniques all offer a more sweeping view of the ocean than can be provided by slow-moving ships and hence have become increasingly important in oceanographic research.

      Satellite-borne radar altimeters have proved to be especially useful. A radar system of this type can determine the distance between the satellite and the sea surface to an accuracy of better than 10 centimetres by measuring the time it takes for a transmitted pulse of radio energy to travel to the surface and return. By combining such a precise distance measurement with information about the satellite's orbit, oceanographers are able to produce maps of sea-surface topography. Moreover, they can deduce the pressure field of the sea surface by combining the distance measurement with knowledge about the geoid. They can in turn extrapolate information about the general circulation of the upper stratum of the ocean from a synoptic view of the surface pressure field.

      Another remote-sensing technique involves the use of satellite-borne infrared and microwave radiometers (radiometer) to measure radiant energy released from the surface of the ocean. Such measurements are used to determine sea-surface temperature. High-resolution, infrared images transmitted by polar-orbiting satellites have provided researchers with an effective means of monitoring wave features in ocean currents over a wide area, as, for example, long equatorial waves in the Pacific Ocean and time variations in the flow of the Gulf Stream between Florida and Cape Hatteras, North Carolina.

      Acoustic (acoustics) techniques also have many applications in the study of the ocean, particularly of those subsurface processes and physical properties inaccessible to satellite observation. In one such technique, the temperature structure of a water column from a given point on the seafloor to the surface is studied using an inverted echo sounder. This instrument, which features both an acoustic transmitter and a receiver, measures the time taken by a pulse of sound to travel from the sea bottom to the surface and back again. In most cases, a change in the average temperature of the water column above the instrument causes a fluctuation in the time interval between the transmission and the reception of the acoustic signal.

      Other acoustic techniques can be utilized to study ocean variables on a large scale. A method known as ocean acoustic tomography, for example, monitors the travel time of sound pulses with an array of echo-sounding systems. In general, the amount of data collected is directly proportional to the product of the number of transmitters and receivers, so that much information on averaged oceanic properties can be gathered within a short period of time at relatively low cost.

Collection of biological samples
      Life at the bottom, benthos, is affected by the water column and by the sediment–water interface; the swimmers, or nekton, are influenced by the water that they come in contact with; and the floaters, or plankton— phytoplankton (plant forms) and zooplankton (animal forms)—are influenced by the water and the transfers that occur at the surfaceofthe sea. Thus, in most cases, measurements and sampling of marine life is best done in concert with measurements of the physical and chemical properties of the ocean and the surface effects of the atmosphere.

      As a consequence of the close interaction of sea life and its environment, marine biologists and biological oceanographers use most of the techniques mentioned above as well as some specialized techniques for biological sampling. Investigative techniques include the use of sampling devices, remote sensing of surface life-forms by satellite and aircraft, and in situ observation of plants and animals in direct interaction with their environment. The latter is becoming increasingly important as biologists recognize the fragility of organisms and the difficulty of obtaining representative samples. The absence of good sampling techniques means that even today little is known about the distribution, number, and life cycles of many of the important species of marine life.

      Some of the most commonly used samplers are plankton nets and midwater trawls. Nets have a mesh size smaller than the plankton under investigation; trawls filter out only the larger forms. The smaller net sizes can be used only when the ship is either stopped or moving ahead slowly; the larger can be used while the ship is travelling at normal speeds. Plankton nets can be used to sample at one or more depths. Qualitative samplers sieve organisms from the water without measuring the volume of water passing through, whereas quantitative samplers measure the volume and hence the concentration of organisms in a unit volume of seawater.

      The Clark-Bumpus sampler is a quantitative type designed to take an uncontaminated sample from any desired depth while simultaneously estimating the filtered volume of seawater. It is equipped with a flow meter that monitors the volume of seawater that passes through the net. A shutter opens and closes on demand from the surface, admitting water and spinning the impeller of the meter while catching the plankton. When the impeller is stopped by closing the shutter, the sampler can be raised without contamination from plankton in the waters above.

      The midwater trawl is specially designed for rapid collection at depths well below the surface and at such a speed that active, fast-swimming fish are unable to escape from the net once caught. Trawls can be towed at speeds up to nine kilometres per hour. To counteract the tendency of an ordinary net to surface behind the towing vessel, a midwater trawl of the Isaacs-Kidd variety uses an inclined-plane surface rigged in front of the net entrance to act as a depressor. The trawl is shaped like an asymmetrical cone with a pentagonal mouth opening and a round closed end. Within the net, additional netting is attached as lining. A steel ring is fastened at the end of the net to maintain shape. A large perforated can is fastened by drawstrings on the end of the net to retain the sample in relatively undamaged condition.

      The use of acoustics to record and measure the distribution of biological organisms is becoming a widely adopted practice. Some organisms can be tracked directly by their distinctive sounds. By recording and analyzing these sounds, biologists are able to chart the behaviour and distribution of such life-forms.

      Organisms that passively affect various electronic systems are large mammals, schools of fish, and plankton that either scatter sound and so appear as false targets or background reverberation, or that attenuate the acoustic signal. Some fishes and invertebrates make up layers of acoustic-scattering material, which may exhibit daily vertical movement related to daily changes in light.

      Light in the upper layers of the ocean is crucial to maintaining marine life. The penetration and absorption of light and the colour and transparency of the ocean water are indicative of biological activity and of suspended material. In situ measurements of water transparency and absorption include the submarine photometer, the hydrophotometer, and the Secchi disk. The submarine photometer records directly to depths of about 150 metres the infrared, visible, and ultraviolet portions of the spectrum. The hydrophotometer has a self-contained light source that allows greater latitude in observation because it can be used at any time of night or day and measures finer gradations of transparency. The Secchi disk, designed to measure water transparency, is a circular white disk that is lowered on a cable into the sea. In practice, the depth at which it is barely visible is noted. The greater the depth reading, the more transparent is the water.

      The primary productivity of the ocean, which occurs in the upper layers, can be monitored by continuous measurement of absorption by chlorophyll molecules. This occurs in the red and blue portions of the spectrum, leaving the green to represent the characteristic colour of biological activity. Satellite measurements of ocean colour that span a number of wavelengths in the visible and infrared portions of the spectrum are used to give a large-scale view of the biological activity and suspended material in the ocean.

Exploration of the seafloor and the Earth's crust
      The ocean floor has the same general character as the land areas of the world: mountains, plains, channels, canyons, exposed rocks, and sediment-covered areas. The lack of weathering and erosion in most areas, however, allows geological processes to be seen more clearly on the seafloor than on land. Undisturbed sediments, for example, contain a historical record of past climates and the state of the ocean, which has enabled geologists to find a close relation between past climates and the variation of the distance of the Earth from the Sun (the Milankovich effect).

      Because electromagnetic radiation cannot penetrate any significant distance into the sea, the oceanographer uses acoustic signals, explosives, and earthquakes, as well as gravity and magnetic fields, to probe the seafloor and the structure beneath. Such techniques—which now include the capability to produce a swath, or two-dimensional, description of the seafloor beneath a ship—are providing increasingly accurate data on the shape of the ocean, its roughness, and the structure beneath. Satellite techniques are a more recent development. Because the shape of the sea surface is closely related to that of the seafloor due to gravity, satellite measurements of surface topography have been used to provide a global view of the ocean bottom. They also have provided data for an accurate mapping of such features as seamounts.

      Research on marine sedimentation (marine sediment) involves the study of deposition, composition, and classification of organic and inorganic materials found on the seafloor. Samples of such materials are thoroughly examined aboard research vessels or in shore-based laboratories, where investigators analyze the size and shape of constituent particles, determine chemical properties such as pH, and identify and categorize the minerals and organisms present. From thousands of reported classifications and collected samples, bottom-sediment charts are prepared.

      Various kinds of equipment are used to obtain samples from the seafloor. These include grabbing devices, dredges, and coring devices.

      Grabbing devices, commonly known as snappers, vary widely in size and design. One general class of such devices is the clamshell snapper, which is used to obtain small samples of the superficial layers of bottom sediments. Clamshell snappers come in two basic varieties. One measures 76 centimetres in length, weighs roughly 27 kilograms (one kilogram = 2.2 pounds), and is constructed of stainless steel. The jaws of this device are closed by heavy arms, which are actuated by a strong spring and lead weight. It is capable of trapping about a pint of bottom material. The second type of clamshell snapper is appreciably smaller. Commonly called the mud snapper, this device is approximately 28 centimetres long and weighs 1.4 kilograms. Other grabbing devices include the orange peel bucket sampler, which is used for collecting bottom materials in shallow waters. A small hook attached to the end of the lowering wire supports the sampler as it is lowered and also holds the jaws open. When contact is made with the bottom, the sampler jaws sink into the sediment and the wire tension is released, allowing the hook to swing free of the sampler. Upon hoisting, the wire takes a strain on the closing line, which closes the jaws and traps a sample. The underway bottom sampler, or scoopfish, is designed to sample rapidly without stopping the ship. It is lowered to depths less than 200 metres from a ship moving at speeds no more than 28 kilometres per hour. The sampler weighs five kilograms and can capture samples ranging from mud to coral.

      The second major category of bottom sampler is the dredge, which is dragged along the seafloor to collect materials. Bottom-dredging operations require very sturdy gear, particularly when dredging for rock samples. A typical dredge is constructed of steel plate and is 30 centimetres deep, 60 centimetres wide, and 90 centimetres long. The forward end is open, but the aft end has a heavy grill of round steel bars that is designed to retain large rock samples. When finer sized material is sought, a screen of heavy hardware cloth is placed over the grill.

      Coring (core sampling) devices typically have three principal components: interchangeable core tubes, a main body of streamlined lead weights, and a tailfin assembly that directs the corer in a vertical line to the ocean bottom. The amount of sediment collected depends on the length of the corer, the size of the main weight, and the penetrability of the bottom. One type of coring device, the lightweight Phleger corer, takes samples only of the upper layer of the ocean bottom to a depth of about one metre. Deeper cores are taken by the piston corer. In this device, a closely fitted piston attached to the end of the lowering cable is installed inside the coring tube. When the coring tube is driven into the ocean floor, friction exerts a downward pull on the core sample. The hydrostatic pressure on the ocean bottom, however, exerts an upward pressure on the core that will work against a vacuum being created between the piston and the top of the core. The piston, in effect, provides a suction that overcomes the frictional forces acting between the sediment sample and the inside of the coring tube. The complete assembly of a typical piston core weighs about 180 kilograms and can be used to obtain samples as long as 20 metres. An improved version of this device, the hydraulic piston corer, is used by deep-sea drilling ships such as the “JOIDES Resolution.” Essentially undisturbed cores of lengths up to 200 metres have been obtained with this type of corer.

      Investigators ay also make use of wire-line logging tools that are capable of measuring electrical resistance, acoustic properties, and magnetic and gravitational effects in the holes drilled. The “JOIDES Resolution” is equipped with tools of this sort, including a remote television camera, which are lowered into a drill hole after the core has been removed. Such wire-line logging apparatus make data immediately available for scientific analysis and decision making.

      Acoustic techniques have reached a high level of sophistication for geological and geophysical studies. Such multifrequency techniques as those that employ Seabeam and Gloria (Geological Long-Range Inclined Asdic) permit mapping two-dimensional swaths with great accuracy from a single ship. These methods are widely used to ascertain the major features of the seafloor. The Gloria system, for example, can produce a picture of the morphology of a region at a rate of up to 1,000 square kilometres per hour. Techniques of this kind are employed in conjunction with seismic reflection techniques, which involve the use of multichannel receiving arrays to detect sound waves triggered by explosive shots (e.g., dynamite blasts) that are reflected off of interfaces separating rocks of different physical properties. Such techniques make it possible to measure the structure of the Earth's crust deep below the seafloor.

Andreas B. Rechnitzer D. James Baker

Additional Reading
The classical tools of the oceanographer are described by H.U. Sverdrup, Martin W. Johnson, and Richard H. Fleming, The Oceans (1942, reissued 1970). More recent texts on oceanography and marine biology include M. Grant Gross, Oceanography, a View of the Earth, 5th ed. (1990); George L. Pickard and William J. Emery, Descriptive Physical Oceanography, 5th enlarged ed. (1990); James P. Kennett, Marine Geology (1982); Bruce A. Warren and Carl Wunsch (eds.), Evolution of Physical Oceanography (1981); James L. Sumich, An Introduction to the Biology of Marine Life, 5th ed. (1992); M.N. Hill et al. (eds.), The Sea (1962– ); and J.P. Riley and G. Skirrow (eds.), Chemical Oceanography, 2nd ed., 8 vol. (1975–83).Exploration and discovery are chronicled in Rachel L. Carson, The Sea Around Us, special ed. (1989); James Dugan and Richard Vahan (eds.), Men Under Water (1965); and Harris B. Stewart, Jr., Deep Challenge (1966). Also informative is the documentary of famous explorers, such as the books by William Beebe, Half Mile Down (1934, reissued 1951); Jacques Piccard and Robert S. Dietz, Seven Miles Down (1961); George Stephen Ritchie, Challenger: The Life of a Survey Ship (1957); Helen Raitt, Exploring the Deep Pacific, 2nd ed. (1964); Willard Bascom, A Hole in the Bottom of the Sea: The Story of the Mohole Project (1961); Francis P. Shepard, The Earth Beneath the Sea, rev. ed. (1967); and Daniel Behrman, The New World of the Oceans (1969). Information on exploration and survey vessels in operation today may be found in an annually updated loose-leaf catalog, Oceanography Vessels of the World (1961– ), published by the U.S. Naval Oceanographic Office. Stewart B. Nelson, Oceanographic Ships, Fore and Aft (1971, reprinted 1983), provides a complete history of U.S. vessels up to the time of publication. See also Jacques Cousteau and Alexis Siverine, Jacques Cousteau's Calypso (1983; originally published in French, 1978); R. Frank Busby, Manned Submersibles (1976), on their design and operation; and Kenneth J. Hsü, Challenger at Sea, trans. from German (1992).Modern undersea explorers are strongly dependent on innovative engineering and technology, such as that treated in Jane's Ocean Technology, 1979–80 (1979); John J. Myers (ed.), Handbook of Ocean and Underwater Engineering (1969); John F. Brahtz (ed.), Ocean Engineering (1968); and Robert L. Wiegel, Oceanographical Engineering (1964). See also John Brackett Hersey (ed.), Deep-Sea Photography (1967); J.F.R. Gower (ed.), Oceanography from Space (1981); G.A. Maul, Introduction to Satellite Oceanography (1985); and F. Dobson, L. Hasse, and R. Davis (eds.), Air-Sea Interaction: Instruments and Methods (1980).Andreas B. Rechnitzer D. James Baker Ed.

* * *


Universalium. 2010.

Игры ⚽ Поможем написать реферат

Look at other dictionaries:

  • exploration — /ek spleuh ray sheuhn/, n. 1. an act or instance of exploring or investigating; examination. 2. the investigation of unknown regions. [1535 45; < L exploration (s. of exploratio) an examination, equiv. to explorat(us) searched out, examined (ptp …   Universalium

  • undersea — [[t]ʌ̱ndə(r)siː[/t]] ADJ: ADJ n Undersea things or activities exist or happen below the surface of the sea. ...an undersea pipeline running to Europe. ...undersea exploration …   English dictionary

  • undersea — un|der|sea [ˈʌndəsi: US ər ] adj [only before noun] happening or existing below the surface of the sea ▪ undersea exploration …   Dictionary of contemporary English

  • undersea — adjective (only before noun) happening or existing below the surface of the sea: undersea exploration …   Longman dictionary of contemporary English

  • Exploration — is the act of searching or traveling for the purpose of discovery, e.g. of unknown people, including space (space exploration), for oil, gas, coal, ores, caves, water (Mineral exploration or prospecting), or information.Although exploration has… …   Wikipedia

  • Earth exploration — Introduction   the investigation of the surface of the Earth and of its interior.  By the beginning of the 20th century most of the Earth s surface had been explored, at least superficially, except for the Arctic and Antarctic regions. Today the… …   Universalium

  • European exploration of Australia — Exploration by Europeans till 1812   1606 Willem Janszoon …   Wikipedia

  • Japan–Korea Undersea Tunnel — The proposed routes for the Japan–Korea undersea tunnel. The Japan–Korea Undersea Tunnel (also: Korea–Japan Undersea Tunnel, JPN–KOR Tunnel, 한일 해저 터널, and 日韓トンネル) is a proposed tunnel project to connect Japan with Republic of Korea (South Korea)… …   Wikipedia

  • space exploration — Investigation of the universe beyond Earth s atmosphere by means of manned and unmanned spacecraft. Study of the use of rockets for spaceflight began early in the 20th century. Germany s research on rocket propulsion in the 1930s led to… …   Universalium

  • Deep-sea exploration — is the investigation of physical, chemical, and biological conditions on the sea bed, for scientific or commercial purposes. Deep sea exploration is considered as a relatively recent human activity compared to the other areas of geophysical… …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”