THE STRUCTURE AND CIRCULATION OF DEEP AND BOTTOM WATERS IN THE ANTARCTIC OCEAN Jeffrey Edwin Callahan THE STRUCTURE AND CIRCULATION OF DEEP AND BOTTOM WATERS IN THE ANTARCTIC OCEAN by Jeffrey Edwin Callahan A dissertation submitted to The Johns Hopkins University in conformity with the requirements for the degree of Doctor of Philosophy. Baltimore, Maryland 1971 • T /4c ?yy LIBRARY KATAL POSTGRADUATE S65008 MONTEREY, CALIF. 93940 ABSTRACT The circulation of deep and bottom waters in the Antarctic Ocean is studied "by analyzing the distributions of heat, salt, and dis- solved oxygen in the ocean south of U0°S. The discussion of the deep-water flow is based on maps showing the depth, salinity, and dissolved-oxygen content of the surfaces where potential steric anomaly equals 50 and 30 centiliters per ton (potential density equals 27-60 and 27.81 grams per liter). The 50-cl/t surface is close to the vertical minimum of dissolved oxygen in the antarctic. This feature is shown to originate in the east- ern South Pacific Ocean. The 30-cl/t surface nearly coincides with the deep salinity maximum, which originates in the North Atlantic Ocean. Maps showing the potential temperature, salinity, and dis- solved oxygen in the bottom layer (depth > 3500 m) of the Antarc- tic Ocean are also included in the report. Bottom water is shown to be produced in both the Ross and Weddell seas. Ross Bottom Water is characterized by high salinities relative to Weddell Bottom Water. The distributions in the bottom layer indicate that the Weddell Sea is a much stronger source of bottom water than is the Ross Sea. The velocity field south of Australia is examined through geostrophic computations and direct current measurements. The eastward Antarctic Circumpolar Current is shown to be divided into two distinct zones separated by weak westward flow. The narrow westward countercurrent is found near the crest of the Indian- Antarctic Ridge. The net eastward flux between Australia and Antarctica is 233 x 10 m3/s, which is nearly equal to the flux through Drake Passage. ACKNOWLEDGEMENTS In preparing this dissertation, I have teen assisted by many people. Foremost among them are Professor Raymond Montgomery of The Johns Hopkins University and Mr. Joseph Reid, Jr. of the Scripps Institution of Oceanography. Professor Montgomery, in his capacity as dissertation advisor, has paid careful attention to this project and has made numerous suggestions for improving the final product. The method of analysis upon which much of the dissertation rests is itself based primarily on the research of Professor Montgomery. Mr. Joseph Reid has, over the past few years, frequently en- couraged and assisted me in my study of oceanography. The oppor- tunities which he provided to participate in oceanographic cruises in the antarctic have had an obvious impact on my choice of re- search topics. I have greatly enjoyed my associations with both of these gentlemen. I have benefited from discussions with members of The Johns Hopkins University Department of Earth and Planetary Sciences. The advice of Professors Owen Phillips and Francis Bretherton and of Associate Professors James Carpenter and Blair Kinsman (now at the University of Delaware) has been particularly helpful. I have received expert technical assistance from the staff of the Chesapeake Bay Institute. Mr. Michael Karweit and Mrs. Carol Feister helped write the computer programs required for data processing. Mr. William Wilson and Mrs. Dean Pendleton pre- ACKNOWLEDGEMENTS ( continued) pared the many figures included in the dissertation. Mrs. Arlene Sullivan has done all the typing. I am grateful to the United States Navy for financial support provided under the Burke Scholar Program. The National Science Foundation supported my work on Eltanin Cruise Ul. Finally, I wish to record my gratitude to my wife for the steady support she has given me during my studies . TABLE OF CONTENTS Page I. INTRODUCTION 1 A. Purpose and Significance 1 B. Previous Work 3 C. Method of Analysis . 12 D. Data Processing and Quality Control 19 II. THE 50-cl/t SURFACE 27 A. Depth 28 B. Salinity 33 C. Dissolved Oxygen 37 III. THE 30-cl/t SURFACE Ul A. Depth Ul B. Salinity 1+1+ C. Dissolved Oxygen 1+9 IV. THE BOTTOM WATER 51 A. Potential Temperature 51 B. Salinity 57 C. Dissolved Oxygen 63 V. FLUX AND VELOCITY STRUCTURE OF THE ANTARCTIC CIRCUMPOLAR CURRENT 66 VI. CONCLUSIONS AND REMARKS 79 FIGURES 86 REFERENCES 122 VITA TABLE OF FIGURES Page Figure 1. Major topographic features of the Antarctic Ocean 87 Figure 2. Stations for the 50- and 30-cl/t maps 88 Figure 3. The 50-cl/t surface: depth 89 Figure h. The 50-cl/t surface: salinity 90 Figure 5. The 50-cl/t surface: dissolved oxygen 91 Figure 6. The 30-cl/t surface: depth 92 Figure 7. The 30-cl/t surface: salinity 93 Figure 8. The 30-cl/t surface: dissolved oxygen 9^ Figure 9- Stations for the bottom-water maps 95 Figure 10. The bottom water: potential temperature 96 Figure 11. The bottom water: salinity 97 Figure 12. The bottom water: dissolved oxygen 98 Figure 13. Drake Passage: salinity section 100 Figure lU. Drake Passage: dissolved-oxygen section 101 Figure 15- Greenwich Meridian: potential-temperature section 102 Figure 16. Greenwich Meridian: dissolved-oxygen section 103 Figure 17. 20 °E: salinity section 10*+ Figure 18. 132°E: salinity section 105 Figure 19. 132°E: dissolved-oxygen section 10o Figure 20. 170°W: salinity section 107 Figure 21. 170°W: dissolved-oxygen section 108 Figure 22. 65°S: salinity section . 109 TABLE OF FIGURES (continued) Page Figure 23. 65°S: dissolved- oxygen section 110 Figure 2k. 132°E: zonal component of relative geostrophic velocity 111 Figure 25. 132°E: zonal component of absolute geostrophic velocity 112 Figure 26. 115°E: zonal component of relative geostrophic velocity 113 Figure 2J . 117°E: zonal component of relative geostrophic velocity 11^ Figure 28. 128°E: zonal component of relative geostrophic velocity 115 Figure 29. lU0°E: zonal component of relative geostrophic velocity 116 Figure 30. lh6°E: zonal component of relative geostrophic velocity 117 Figure 31. 20°E: zonal component of relative geostrophic velocity 11° Figure 32. Discovery dissolved-oxygen correction 119 Figure 33. Station curves for the Ross Sea and the Weddell Sea 120 I. INTRODUCTION I. A. Purpose and Significance The Antarctic or Southern Ocean, which I will take to be the re- gion between U0°S and the Antarctic continent, plays a central role in the deep and bottom circulation of the world ocean. There are two fundamental reasons for its importance in these phenomena: (l) The Antarctic Ocean is by far the largest con- nection among the Atlantic, Indian, and Pacific oceans. Enormous amounts of water, heat, salt, and nutrients are exchanged among the oceans. The exchange helps maintain the observed tempera- ture, salinity, and dissolved-oxygen fields in the world ocean, thereby affecting the earth's heat budget and the distributions of marine organisms. (2) The extreme southern reaches of the Antarctic Ocean are among the few regions of the world where climatic conditions may lead to navifacial1 water becoming suf- ficiently dense to overturn and replenish the deep and bottom water. These sources of bottom water are critical elements in the thermohaline circulation of the ocean. Insight into the various dynamic and thermodynamic processes which determine the antarctic circulation may be gained from tak- ing several points of view. Theoretical attacks on the problem l The term "naviface" (Montgomery, 19&9) will be used to distin- guish the air-sea interface from the several other surfaces discussed in this paper. have so far centered on modeling the quasi-zonal Antarctic Cir- cumpolar Current, which is the dominant flow in the region. These studies have been reviewed in the author's masters essay (1968, unpublished). The present paper takes a different approach, an essentially empirical one. The horizontal and vertical distributions of temperature, salinity, and dissolved oxygen in the abyssal lay- ers are examined in detail. Within the limitations of certain simplifying assumptions, to be discussed later, conclusions are then reached concerning the large-scale structure of the circu- lation. Similar studies have been undertaken previously. Most of them were based on data gathered prior to World War II. Within the last decade a large amount of hydrographic work has been carried out in the antarctic. Much of this work has been done in areas that were previously very poorly sampled (e.g.3 the Weddell Sea) or in areas where old measuring techniques were not sufficiently sensitive to detect differences in water character- istics [e.g., the Pacific Ocean). Further, the recent stations include samples taken closer to the sea floor than on the early stations. Although the new data have been discussed in relation to specific sections of antarctic waters, notably by A. L. Gordon, the present paper represents the first attempt to incorporate these data into a general analysis of the circumpolar region. In part, this study was prompted by the availability of a large body of new data and a desire to find out what new features might "be revealed in the deep and bottom waters of the antarctic. The investigation was also motivated by an interest in some spe- cific aspects of the antarctic circulation which are not yet well understood. These include the structure of the velocity field and the transport of the Antarctic Circumpolar Current, the merid- ional exchange of deep water between the antarctic and the oceans to the north, and the importance of the Ross Sea as a source of bottom water. I.B. Previous Work Coir present understanding of the large-scale structure and cir- culation of antarctic waters has been pieced together from the efforts of many individuals . Most of this work has been carried out within the last forty years, and my review will therefore focus on this period. However, it is interesting to note that at least a partial appreciation of the role of the antarctic in the deep oceanic circulation was evident somewhat earlier. Buchan (1895, p. 8) concluded from the deep-sea temperature measurements made by the Challenger (1872-76) that, "There can be no doubt that these very low deep-sea temperatures have their origin in the Southern or Antarctic Ocean, the icy-cold waters of which are propagated northwards... ." Buchan (p. 27) also deduced the ex- istence of a warm deep current flowing from the North Atlantic to the South Atlantic. This water mass, now known as North Atlantis Deep Water, is another major component of the abyssal circulation in the antarctic. During the 1920' s, subs t ant ial improvements were made in the hydrographic coverage of the South Atlantic, Indian, and Pacific oceans. For the first time, oceanographers could begin to understand the character of the general circulation. On the basis of Carnegie (1928-29) stations taken in the Pacific Ocean, Sverdrup (1931) was able to show that the deep circulation of the Pacific must be qualitatively different from that of the Atlantic and Indian oceans. He concluded that the Pacific deep water is a mixture of deep water from the Atlantic and Indian oceans with bottom water from the antarctic. Owing to a scarcity of data, Sverdrup was unable to venture beyond these general conclusions, and some of the details of the flow he deduced were later proved incorrect. Nonetheless, the concept of inter-ocean exchange which he formulated is basic to our present understanding of the abyssal circulation. In 1933, three papers dealing with the characteristics and circulation of Atlantic -Ant arctic waters were published. Wiist (1933) discussed at length the distributions of potential temper- ature and salinity near the sea floor in water deeper than U000 m. His maps, which were based on all data available at that time, showed in a very striking fashion the propagation of cold bottom water out of the Weddell Sea, northward along the western trough of the Atlantic and eastward into the Indian Ocean. Wust's maps have become the classic piece of evidence for the importance of the Weddell Sea as a source of bottom water, although there is some ambiguity in Wust's salinity map. A more general description of the hydrography of the western Atlantic-Antarctic was written by Deacon (1933). His study was based on the early cruises of the Discovery and Disoovevy II. In a brief section on the bottom water, Deacon reviewed several pro- posed mechanisms for bottom-water formation. He concluded that the bottom water is probably formed over the continental shelf of the Weddell Sea, as Brennecke (1921, referenced by Deacon, 1933) had proposed somewhat earlier. The question of exchange between the Atlantic and Pacific oceans via Drake Passage was examined by Clowes (1933) in his study of the deep water of the Scotia Sea. Noting that the deep water of the Scotia Sea is somewhat less saline than that farther to the east, he concluded that most of this water must come from the Pacific. Clowes computed the geostrophic transport relative to 3500 db through Drake Passage and arrived at a net eastward flux of 110 x 106 m3/s. This was the first quantitative estimate of the exchange of water through Drake Passage. A broad account of the hydrography of the Atlantic-Antarctic sector south of 55°S, between U0°W and U0°E, was given by Mosby (193*0. A large part of his data were from the Norvegia Expedi- tions of 1927-28, 1929-30, and 1930-31, but Deutschland (1911-12), Meteor (1925-26), and William Sooresby (1931) stations were also included. In addition to describing and classifying the water masses of the region, much as Deacon (1933) had done, Mosby dis- cussed two dynamical features of interest: (1) His maps of dynamic topography (e.g.3 his Figure 28) show clearly a weak, elongated cyclonic gyre centered in the northeastern Weddell Sea and extending as far as 15°E. Moshy pointed out that this flow pattern is consistent with the drift of the Deutschlandj which was icebound in the Weddell Sea during the winter of 1912. Noting that the winds in this region also have a roughly cyclonic pattern, shifting from westerlies north of 60°S to easterlies along the antarctic coast, he concluded that the Weddell gyre is driven by the wind. The existence of a cyclonic gyre in the Weddell Sea is also suggested by Mosby's map of maximum sub-surface temperatures, on which a westward intrusion of relatively warm water along the antarctic coast can be seen. According to Mosby, this feature had been observed in the Deutsehland data by Brennecke (1921), who theorized that the warm water comes directly from the Indian Ocean. Mosby, however, recognized the warm water as the uniform deep water found elsewhere around Antarctica, and ruled out a direct Indian Ocean influence. (2) Mosby devoted considerable space to the question of bottom-water formation. After reviewing the theories of bottom- water formation, he, like Deacon, concluded that Brennecke' s (1921) explanation is correct. Brennecke had proposed that bottom water is formed in the Weddell Sea from navifacial water which is overturned during winter. The postulated vertical convection, which takes place in shallow water, is driven by cooling and the addition of salt at the naviface during the formation of sea ice. Dense, uniform shelf water then flows down the slope, mixes with warmer, saltier deep water, and forms a dense mixture which sinks to the bottom. In accepting this mechanism for bottom-water formation, Mosby pointed out that to produce bottom water with the observed characteristics, the shelf water (assumed to be at the freezing point) must have a salinity of at. least 3^.62/?c . Mosby rejected Wust's (1933) suggestion that the mechanism proposed by Brennecke is supplemented by deep vertical convection occurring "at the edge of the convergence area created by pack ice..." (Wust, 1933, p. Vf) • Wust's theory seemed highly unlikely because (l) no such convergence has been detected (in fact, the prevailing wind field would suggest the existence of a region of divergence), and (2) the observed vertical temperature and salin- ity distributions are inconsistent with deep overturning. The first study of the circumpolar characteristics of the Antarctic Ocean was by Deacon (1937) • The principal source of data for Deacon's work was the circumpolar cruise of the Dis- covery II in 1932-33, during which nineteen sections were made in the antarctic zone. Several highly significant features of the deep and bottom flow around Antarctica were revealed in Deacon's paper. It be- came clear for the first time that bottom water is not formed all around the antarctic coast, as proposed by Sverdrup (1931), but is instead formed primarily in the Weddell Sea. Deacon also showed that the bottom water formed in the Weddell Sea flows 8 around Antarctica towards the east, not, as Sverdrup had deduced, towards the west. (These conclusions were later confirmed by Sverdrup (19AO) on the basis of data taken in the Indian -Ant arctic Ocean during the BANZAR Expedition2 ( 1929-31) . ) Another valuable result of the Discovery data was to confirm the extreme uniform- ity of the deep water around Antarctica, which Sverdrup (1931) had predicted from the limited data available to him. In his BANZARE Report Sverdrup acknowledged the defects in his earlier theory of the deep circulation, and he proposed a revised version. He again emphasized the importance of the Atlantic Ocean as the primary source of deep water in the world ocean and the role of the Antarctic Circumpolar Current as the agent by which Atlantic deep water is transported to the Indian and Pacific oceans. Sverdrup computed the eastward water trans- port relative to 3000 db across a section from Tasmania to Antarctica (lU5°E). His result, ikk x 106 m3/s, was the first estimate of the flux of water from the Indian to the Pacific ocean. Another significant aspect of the circumpolar circulation was pointed out by Sverdrup in his discussion of the general cir- culation in The Oceans (Sverdrup et al.3 19^2) . Sverdrup noted a systematic deflection of the predominantly zonal transport lines by major topographic features (see his Figure 163) . The direction 2 British, Australian & New Zealand Antarctic Research Expedition of the deflection (equatorward on approaching a ridge; poleward after crossing over the ridge) is in agreement with qualitative predictions made by assuming conservation of potential vorticity. One implication of these deflections is that the Circumpolar Cur- rent must extend to great depths. During the period from World War II through the mid-1950's little oceanographic research was carried out in the antarctic, and few papers were published on the subject. One paper from this period which is notable in the present context is Fofonoff 's (1956) study of bottom-water formation. Using temperature and salinity data collected in the northern Weddell Sea, Fofonoff gave quantitative support to Brennecke's model of bottom-water formation. The International Geophysical Year (1957-58) brought about a resurgence in all phases of antarctic research including ocean- ography. Over the past decade, research ships from the United States and the Soviet Union have been very active in the antarc- tic, and many hydrographic stations have been occupied. The sub- stantial increase in field activity has resulted in numerous pub- lished reports on antarctic oceanography. Ishino (1963) presented circumpolar plots of temperature, salinity, and density at several depths from 50 m to 3000 m, using mainly Discovery (1925-37) and Ob ' (1956-58) stations. He attempted to discuss seasonal variations in characteristics down to 2000 m, but this effort was severely limited by the lack of winter data. Perhaps the strongest conclusion that may be drawn 10 from Ishino's maps is that no systematic seasonal variations in temperature and salinity are evident below 200 m, at least in those areas where sufficient data are available for comparison. Ishino (p. 166) apparently believed that large amounts of bottom water are produced in shallow regions along the Indian Ocean sec- tor of the antarctic coast. This conclusion conflicts with the evidence presented by Sverdrup (19^+0), Deacon (1937) » and the present work. The most recent descriptive studies of the Antarctic Ocean are those by Gordon (1966, 1967b, 1970) and Reid and Nowlin (1971). The first two papers by Gordon are analyses of the sector from 20°W to 170 °W, utilizing the early Eltanin data. Gordon (1970) has extended his study of bottom water into the eastern Indian Ocean. Gordon and Goldberg (1970) have prepared a folio contain- ing maps of temperature, salinity, and dissolved oxygen at con- stant depths and sections showing the general structure of ant- arctic waters. Pertinent aspects of these studies will be dis- cussed in later sections. Reid and Nowlin (1971) discuss the deep circulation in the Drake Passage region. By using deep current measurements made in the Passage as reference values for geostrophic computations, they have determined that the net eastward mass flux through Drake Passage is well over 200 x 106 m3/s (their value is 237 x 106 m3/s) Most previous estimates were on the order of 100 x 106 m3/s. The results of the above studies may be summarized in the following brief description of deep and bottom waters in the 11 antarctic: Circumpolar Deep Water (Sverdrup, 19^+0) is a thick, uniform layer of relatively warm and salty water found in all longitudes around Antarctica. It has the following characteris- tics: 8 = 1.0 - 2.0°C, S = 3^.70 - 3U.76 t ,0 = U.O - U.6 ml/1. 2 Circumpolar Deep Water flows in a broad, quasi-zonal current east- ward around the continent at speeds of roughly 5 cm/s , and partic- ipates in a much slower meridional exchange with the oceans to the north. The mean zonal transport of the Antarctic Circumpolar Cur- rent is probably in excess of 200 x 106 m3/s. Permanent large- scale meanders in the zonal flow are found over major topographic features . A major constituent of Circumpolar Deep Water is North At- lantic Deep Water, which moves south and mixes into the circumpolar flow in the Atlantic and Indian ocean. The magnitude of the south- ward flux of North Atlantic Deep Water is difficult to determine, but calculations by Wright (1969) and Sverdrup et at. (19^2) in- dicate that it is on the order of 20 x 106 m3/s at 30°S. As the Circumpolar Deep Water flows around Antarctica, its properties are further modified by lateral and vertical mixing, ultimately producing the extremely uniform deep water (Common Water, Mont- gomery, 1958) of the Pacific Ocean. The second major water mass found in the abyssal depths of the antarctic zone is Antarctic Bottom Water. (For the present, Antarctic Bottom Water will be referred to as a single water mass. Later I will show that two distinctive water types can be recog- nized within this mass.) Antarctic Bottom Water has the char- 12 acteristics: 9 = -0.8 - 0°C, S = 3^.65 - 3U.71 %* , 0 = 5.U - 2 7.0 ml/1. The primary source of bottom water is the Weddell Sea, al- though the Ross Sea probably does produce limited quantities (Gordon, 1966, 1970). The bottom water formed in the Weddell Sea is a mixture of very cold, moderately salty shelf water and warmer, more saline deep water. The dense shelf water is supposed to be produced during winter when sea ice forms. This mechanism is only inferred from summer observations , since no observations of shelf water have ever been made in winter. I.C. Method of Analysis In this paper the circulation is studied by means of maps and sections showing the distributions of salt and dissolved oxygen in the deep and bottom waters . These substances act as tracers with which the mean flow field can be deduced using a model based on the diffusion equation. Several assumptions and approximations have been made in order to simplify the interpretation of the maps. The first of these is that the large-scale ( > 102km) distributions of heat, salt, and dissolved oxygen are in a steady state below the shallow navifacial layer. In view of the demonstrated reproducibility of temperature and salinity measurements below the naviface (compare Wiist , 1936, Teil A and Fuglister, i960) this appears to be a reasonable approx- imation, although the period over which such measurements have been 13 made is admittedly short in comparison with the residence time of water in the deep ocean. Both salt and dissolved oxygen are treated as conservative properties, meaning that there are no sources or sinks below the naviface. In the case of dissolved oxygen this is obviously not strictly true. Phytoplankton concentrated in the upper 100-200 m provide a source of oxygen, and decaying organic matter through- out the water column acts as a distributed sink. Below the euphotic zone the net effect is equivalent to a weak sink, the strength of which is not well known. Wright (1969) quoted deep 0 -depletion rates of U.3 - 53.0 x 10 ml/1 per year; data col- 2 lected on the 19&9 GE0SECS test station suggest an average rate of 25 x 10_I+ ml/1 per year (Craig and Weiss, 1970). The time scale of the deep circumpolar flow is roughly 20 years. At an average depletion rate of 25 x 10 ml/1 per year, changes in oxygen-concentration due to the sink term would amount to 0.05 ml/1 over 20 years. This change is small compared with the prob- able error of the oxygen data. The third fundamental assumption behind the analysis is that large-scale flow in the ocean tends to take place along surfaces of uniform potential density or potential steric anomaly. Because of geostrophy, such surfaces often have a small ( < 1:103) slope relative to level {i.e., uniform-depth) surfaces. Over oceanic scales, these small slopes may be significant, and it is there- fore important to distinguish between surfaces of uniform depth and surfaces of uniform potential steric anomaly. Processes oc- lU curring along the latter surfaces will be designated as "lateral" processes {e.g.3 lateral mixing) as opposed to "horizontal" pro- cesses occurring on surfaces of uniform depth. The degree to which the above assumption is valid cannot be stated quantitatively. If there were no diffusion or mixing in the ocean, this approximation would necessarily be an exact state- ment, since the potential steric anomaly of a fluid element is de- fined by its salinity and potential temperature, which are conser- vative properties in the sense stated previously. It does seem reasonable to assume that in a stratified system (such as the ocean) fluid elements are constrained somewhat to flow along, rather than across, surfaces of uniform potential steric anomaly. Ultimately, however, use of the assumption rests on empirical grounds {e.g., Reid, 19&5, Figure 3) and on the proven value of this model. As a consequence of the third assumption, vertical advection is no longer a separate term in the diffusion equation, having been absorbed in the lateral advection terms. At this stage the model has been reduced to a balance between lateral advection and diffusion. In the large-scale physical processes of the ocean, molecular diffusion is negligible compared with turbulent diffu- sion, or mixing, which represents the gross effect of the fluc- tuating advective terms on the mean field. Turbulent mixing is customarily decomposed into vertical and lateral components. In studies like the present one it is often assumed (at least tacitly) that vertical mixing is of secondary importance compared 15 with lateral mixing. This assumption is probably due to the heavy emphasis placed on lateral mixing by Rossby (1936, 1937) » Montgomery (1938), and others who developed the method used here. In most cases , such an assumption is not inconsistent with the distributions of properties in the open ocean. On the other hand, even in those studies where lateral mixing is assumed to predom- inate there may be regions where the observed patterns can be ex- plained only in terms of vertical mixing. That is, on the par- ticular surface in question, there is no source of water with the characteristics necessary to produce the observed distributions. Theoretical studies are of little assistance in deciding whether lateral or vertical mixing should predominate. The large- scale oceanic distributions are matched equally well by models containing (l) only vertical mixing (Robinson and Stommel, 1959) » (2) only lateral mixing (Sverdrup, 1939) » or (3) no mixing at all (Welander, 1959)- Some of the confusion is no doubt due to our imperfect understanding of the mechanisms involved in turbulent mixing, and the consequent use of rather arbitrary mathematical functions to represent the mixing terms. In the present work I am not concerned with the analytical form of the mixing terms but only with the qualitative effects of mixing. The relative importance of vertical and lateral mixing is determined from the observed distributions. Except in a few cases where vertical mixing is clearly important, the distributions are assumed to represent a balance between lateral mixing and lateral advection, 16 with the mean flow taking place along the isopleths . Three sets of maps have been drawn for this study. The first two sets show the salinity, concentration of dissolved oxygen, and depth of the surfaces where potential steric anomaly equals 50 and 30 centiliters per ton3 (abbreviated cl/t). Poten- tial steric anomaly, 6Q, is defined by the following equation: (1) 6e = as,e,o " a35,o,o where a „ - = specific volume of sea water of salinity s, potential temperature 8, at sea pres- sure 0 db ; potential temperature is the temperature a sea water sample would have after being raised adiabatically to the naviface (p = 0) ; and a = specific volume of sea water of salinity ->s jU »U ¥i%e, potential temperature 0°C, sea pressure 0 db, = 0.9726U ml/g. Surfaces of uniform potential steric anomaly will be called isan- osteric surfaces . C.-G. Rossby appears to have been the first to use the dis- tribution of a conservative property along a surface of uniform potential density to deduce geophysical circulation patterns. 3 These surfaces correspond to potential density surfaces aQ = 27.60 and 27.81 g/1, respectively. IT The method was applied to the atmosphere by Rossby and collabora- tors (1937) > who used specific humidity as a tracer on surfaces of constant potential temperature (isentropic surfaces). Because of the transient nature of even the large-scale atmospheric motion, isentropic analysis, as it was called, has not found much application in synoptic meteorology. However, the concept has proven valuable in studies of the ocean, where, as I noted earlier, the large-scale circulation is reasonably steady. The earliest oceanographic applications of Rossby' s technique were by Parr (1938) and Montgomery (1938). Parr discussed the distribution of salinity on several surfaces of uniform a ( ^ potential density) south of Newfoundland. Montgomery made a detailed study of the circulation in the tropical North Atlantic using the distributions of salinity and dissolved oxygen on six Q -surf aces . Montgomery chose to call the method isentropic anal- ysis in conformity with meteorological usage. This terminology is still commonly seen. In view of the fact that isanosteric surfaces (or surfaces of uniform O ) are not in general isentropic surfaces, the expression "isanosteric analysis" suggested by Tsuchiya (1968) may be more appropriate. Notable applications of isanosteric analysis have been made in recent years by Taft (19&3), whose maps of salinity and dis- solved oxygen at intermediate depths cover the southern hemisphere; Reid (1965) » who examined the intermediate waters of the Pacific; Tsuchiya (1968), who studied the tropical Pacific; and Barkley 18 (1968), who has prepared an atlas of the Pacific Ocean. All the above studies have used isanosteric surfaces which are confined to the shallow and intermediate layers of the ocean. When surfaces which extend to great depth (as in the present work) are used, an additional complication is introduced. The problem is caused by a rather subtle effect pointed out by Montgomery (1937) some time ago, but generally ignored until quite recently (Lynn and Reid, 1968) . The compressibility of sea water is a function of its temperature, cold water being more compressible than warm. As a consequence, two fluid elements of equal steric anomaly, but different temperatures, at the naviface will have different steric anomalies if both are moved adiabatically to some depth, say 2000 m. In other words, fluid elements which could mix laterally at the naviface could not do so at 2000 m. The implication is that a single isanosteric surface does not define a surface of lateral advection and mixing at all depths. Necessary conditions for the above effect to be important are that the temperature range on the isanosteric surface be significant (2-3°C) and that the isanosteric surface undergo large variations in depth ( > 2000 m) . In the case of the 30- cl/t surface, the depth range is large ( ^ 2500 m) , but the temperature range is small ( < 1°C) over the region where most of the variation in depth occurs. The largest temperature gra- dients are found near Antarctica, where the 30-cl/t surface is shallow and pressure effects are not significant. Similar condi- tions obtain on the 50-cl/t surface. It therefore appears that 19 the errors introduced by using a single isanosteric surface to represent a surface of lateral advection and mixing are not ser- ious in this particular study. In the antarctic, the water "below the 30-cl/t surface is extremely uniform with respect to density. Small measurement errors or interpolation errors lead to substantial errors in computing the characteristics of deeper isanosteric surfaces. Therefore, isanosteric analysis is not useful for studying the water near the bottom. The third set of maps prepared for this paper shows the distributions of potential temperature, salinity, and dissolved oxygen near the sea floor. Data were plotted only in regions where the depth of water exceeds 3500 m. With few exceptions the samples are within 300 m of the bottom; the great majority are within 200 m of the bottom. Deacon (1937) and Wust (1938) have prepared circumpolar maps of bottom potential temperature in depths greater than U000 m. These maps were based on a much sparser data grid than is now available. Gordon (1966) plotted maps of bottom potential temper- ature and dissolved oxygen in depths greater than 3000 m for the sector 20 °W to 170 °W. The bottom salinity and dissolved oxygen maps presented in the present paper are the first such maps to be drawn for the entire circumpolar zone. I.D. Data Processing and Quality Control The bulk of the data used in this study was acquired from the 20 National Oceanographic Data Center (NODC) , Washington, D. C. Additional data have been acquired from various sources shown in Table 1, which lists the data by ship in chronological order. Over three-quarters of the stations were taken by three ships : Eltanin (kl%) 9 Discovery (27%), and Ob' (9%). All but three of the stations are Nansen casts, the exceptions being Eltanin S-T-D stations. Figure 2 shows the geographical distribution of stations for the isanosteric maps, and Figure 9 shows the distribution of stations for the bottom maps. Most, but not all, of the bottom stations were used in the isanosteric maps too. Note the rather sparse coverage of the central Indian Ocean. Approximately 75% of the stations were occupied during the six months November-April, with 50% of the total being occupied in the three months January -Mar ch , the austral summer. The hydrographic data were processed on The Johns Hopkins University Computing Center IBM 709^. Potential temperature and potential steric anomaly were computed using equations from Fofonoff (1962a). Linear interpolation was used to determine the values of parameters on the 50- and 30-cl/t surfaces . All data were plotted and contoured by hand on polar sterographic maps of scale 1:15 x 106. Bottom topography, Figure 1, was taken from U. S. Naval Hydrographic Office charts H.O. l6, 892-1 (1959) and H.O. 1262A and from the Soviet Atlas Antarktiki (Tolstikov, 1966) . In order to avoid crowding on the smaller-scale maps 21 TABLE 1. Sources of data. SHIP NATION TOTAL BOTTOM DATA YEARS STATS. STATS. SOURCE Meteor Germany 1925-26 32 11 1 Discovery Gt. Brit. 1931-39 1951 240 39 1 Thorshaven Norway 1934 1 0 1 Galathea Sweden 1952 1 1 1 Ob' USSR 1956-63 1965-66 78 8 1 Capitan Canepa Argen. 1957-59 1962 10 4 1 Horizon US 1957 5 1 1 San Martin Argen. 1958 1 0 1 Operation Deep Freeze: various Navy & Coast Guard icebrkrs US 1957-61 1964 1967 22 0 1 Atlantis US 1959 6 k 2 Vema US 1959-60 6 3 1 Argo us 1960-61 1962 5 k 1 Diamantina Austral. 1960-62 15 4 1 Gas coyne Austral. 1961 6 0 1 Umitaka-Maru Japan 1961-62 21 0 1 Africana II S. Africa 1962-64 8 0 1 Eltanin us 1962-67 1967 1968 1970 362 208 1 3 4 5 22 TABLE 1. (continued) SHIP NATION YEARS TOTAL BOTTOM DATA STATS. STATS. SOURCE Anton Bruun US 196 k 2 0 1 Glacier (IWSOE) US 1968-69 22 5 6 Hakuho-Maru Japan 1969 18 Ik 7 T. Washington US 1969 21 Ik 8 882 320 1. National Oceanographic Data Center 2. Fuglister (i960) 3. Cruises 28, 29 (Scorpio Expedition) : SIO/WHOI (1969) h. Cruises 32-36: Jacobs et at. (1970) 5. Cruise kl: author (unpublished data) 6. International Weddell Sea Oceanographic Expedition: Hufford & Tennyson (1970) 7. Tokyo University Ocean Research Inst. (1970) 8. Piquero Expedition: SIO (unpublished data) 23 presented here, only station positions have been plotted. Data quality and quality control are extremely important considerations in a study of this type, for they determine the resolution and reliability of the contours. The precision of routine shipboard temperature and salinity measurements has been well established by accumulated experience and by statistical analyses of results (see, for example, Wooster and Taft , 1958). Reversing-thermometer measurements are accurate to better than 0.02°C. Salinities determined by chlorinity titration are gen- erally good to within 0.02 %a , although greater precision can be achieved by taking special precautions, as in the case of Meteor salinities. Salinities measured with electrical conduc- tivity salinometers are significantly more accurate than those done by titration. Cox (1965) states a precision of 0.002 %> f or these instruments. In routine use the reliability is probably somewhat less, perhaps 0.00 if %0 . The reliability of the diss olved-oxy gen determinations is difficult to estimate. Nearly all methods for determining the concentration of dissolved oxygen in sea water are based on the Winkler method. Under optimum conditions this method may achieve an accuracy of 0.1$ (Carpenter, 1965) . However, as Carritt and Carpenter (1966) have demonstrated, numerous variations in ana- lytical technique among oceanographic institutions and individual chemists make it impossible to establish the routine reliability Virtually all the salinity data since 1962 (about half of the total) were measured by this method. 2U of oxygen data. Wide variations in quality may be expected. Of particular significance to this study are the oxygen measurements made during the Disaovery cruises . Several authors (Wooster and Volkmann, i960 ; Gordon, 1966; Carritt and Carpenter, 1966) have noted that these measurements are consistently lover than those made by other ships in the same vicinity. The dis- crepancy has been reported to be on the order of -0.5 ml/1 by Gordon (1966) and Carritt and Carpenter (1966). The present author has made numerous comparisons between the Discovery dissolved-oxygen data and those of other ships at the same depth and approximate geographical location. The consistent difference observed by others was confirmed. The possibility of a long-term increase in dissolved-oxygen concentration was ruled out by the fact that Disaovery values were also lower than Meteor (1925-26) data, and that no trend in oxygen concentration was ob- served in the Discovery data, which were taken over a ten-year period. Unfortunately, the raw Discovery data were destroyed in World War II, and it is not possible to determine and correct the error properly. I decided to devise a correction curve by comparing the measured navifacial oxygen concentration with the saturation value appropriate to the measured navifacial tempera- ture and salinity. The resulting linear best-fit shown in Figure 32 was used to correct all Discovery oxygen data. The correction is proportional to the dissolved oxygen concentration, ranging from +0.2 ml/1 at a measured value of U.O ml/1 to + 0.U ml/1 at 7.0 ml/1. 25 A similar trend (increasing error with concentration) was observed from the comparison of deep Discovery oxygen measure- ments with those of other ships. However, the corrections de- rived by extrapolating those comparisons are somewhat larger than those found from the navifacial -concent ration/saturation- value method. The differences between the two methods might be partly accounted for by the fact that antarctic navifacial waters are commonly supersaturated (Jacobs and Amos, 1967) » leading to an underestimate of the navi face -measurement error. On the whole, it was felt that the navi face-comparison method was a more reli- able correction procedure, since it was based on a wider range of data points. It is likely, however, that the corrected Discovery oxygen data are still, on the average, lower than neighboring points from other ships . Aside from the systematic correction applied to all Discovery dissolved-oxygen data, no corrections were made to the original data. Questionable points were evaluated, and occasionally elim- inated, on the basis of the following objective and subjective criteria: (1) agreement with neighboring points on the same surface, and consistency of vertical profile; (2) method of making measurement; (3) general reliability of data from the particular ship or cruise in question. The "general reliability" test was based on repeated compar- isons of stations from several ships in close proximity to each 26 other. As a rule, these comparisons showed good agreement among the data of different ships. The notable exception (in addition to the Discovery oxygens) was the Ob ' salinity measurements, which contain more scatter than those of other ships. Since most of the Ob * stations used are close to the Antarctic continent, where the lateral gradients are relatively strong, this scatter is not serious . 27 II. THE 50-cl/t SURFACE The 50-cl/t isanosteric surface was chosen for analysis because it forms what might be considered the upper boundary of the deep water mass being studied here. It is recognized that the defini- tion of such a boundary must be somewhat arbitrary, being based on gradients in the characteristics which distinguish each water mass . Perhaps the most fundamental and unambiguous of these char- acteristics is salinity. South of about 55°S, the 50-cl/t sur- face lies within the sharp halocline which marks the transition from the fresh upper waters of the antarctic to the more saline deep water. North of about 50°S (the "Antarctic Convergence"), the 50-cl/t surface is found between the salinity minimum (Ant- arctic Intermediate Water) at 80 cl/t and the deep salinity max- imum (Circumpolar Deep Water) at 30 cl/t (see Figures 18, 20). Over much of the antarctic region (indeed, over a large part of the southern hemisphere) , the 50-cl/t surface coincides with the vertical minimum of dissolved oxygen (see Figures l6, 19, 2l) . This relationship persists from the subtropics to about 60°S, ex- cept in the eastern South Atlantic, where the two coincide between roughly 30°S and 50°S. In this respect the 50-cl/t surface again provides a convenient boundary between the upper and intermediate waters and the deep water, all of which have higher dissolved- oxygen concentrations. 28 II. A. Depth The depth of the 50-cl/t surface (Figure 3) ranges from less than 100 m near Antarctica to more than 1600 m at about U0°S. During the summer season, this isanosteric surface lies below the naviface everywhere except in some parts of the Ross and Weddell seas. The data used in this study show the 50-cl/t sur- face intersecting the naviface at several stations along the western side of the Ross Sea and at two stations very close to the western and southern boundaries of the Weddell Sea. Only two years' (1968, 19&9) summer data are available for the Weddell Sea, and it is not known how representative they are of general conditions there. In the Ross Sea, on the other hand, numerous stations have been occupied by U. S. icebreakers over the past fifteen years. These data show that navifacial steric anomaly values less than 50 cl/t are not uncommon in the southeastern and western portions of the Ross Sea. It appears that the 50-cl/t surface frequently intersects the naviface in the interior of the Ross Sea during summer. In view of the similarity of climatic conditions in the Ross and Weddell seas, it is not unlikely that this situation occurs intermittently in the extreme reaches of the Weddell Sea too. The only winter observations from either of these areas were taken by the Deuts ohland in the Weddell Sea in 1912. These stations show navifacial steric anomaly values well below 50 cl/t (but above 30 cl/t) occurring throughout the central Weddell Sea. 29 Indirect evidence of winter conditions is provided by the very cold water (Winter Water, Mosby, 193^) which is observed just below the thin navifacial mixed layer near the Antarctic contin- ent in summer. This water, which is the remnant of the homoge- neous navifacial layer formed in winter, has the following char- acteristics: 9 < -1.0°C, S ^ 3U.3 X , 6. < 50 cl/t. Therefore, y it is highly probable that in winter the 50-cl/t surface inter- sects the naviface several degrees from the coast all around Antarctica. The depth contours of an isanosteric surface provide a qualitative representation of the large-scale geostrophic circu- lation. The dominant pattern evident in the contours of the 50- cl/t surface is the quasi-zonal, eastward-flowing Antarctic Cir- cumpolar Current. In most longitudes the zonal flow occupies a band of latitude some 10° - 15° in width, as defined by the dis- tance between the 200-m contour and the lUOO-m contour in the Atlantic and the l600-m contour in the Indian and Pacific oceans. However, in several places the flow constricts tightly to a width of about 5°> indicating significantly increased velocities in these areas. Such constrictions occur in Drake Passage, across the Macquarie Rise, and north of the Ross Sea at lU5°W. The first two of these constrictions are obviously forced by contractions of the lateral boundaries of the flow, but the con- striction north of the Ross Sea cannot be explained in terms of lateral boundary effects. Gordon (1967b) has stated that the flow pattern here is caused by the "channelling" effect of the ■ 30 fracture zone (called by Gordon the USARP fracture zone) which cuts across the Pacific-Antarctic Ridge at lUo°W. Although the fracture zone may play some part in constricting and turning the flow north of the Ross Sea, it is difficult to "believe that such a striking change in the flow could be caused by this relatively small topographic feature. In all of the places where constrictions of the flow occur, and over the Scotia Ridge too, wave-like distortions of the zonal flow are observed. These features are in approximately the same positions as those noted by Sverdrup et at. (19^2). Close to the antarctic coast there is a tendency for the 50-cl/t surface to deepen slightly toward the south. The reverse slope, which is most pronounced from 0° to lU0°E, is indicative of westward flow along the coast. Deacon (1937) gave the name "East Wind Drift" to the shallow coastal current. The name suggests the probable cause of the westward flow near the coast: easterly winds prevail south of about 65°S. Barcilon (1966) has shown that such a current might be driven by the discharge of melt water from the antarctic ice cap. Even if this is not the case, significant seasonal variations in the coastal current due to changes in the sea-ice cover and melt-water runoff might be expected. For ob- vious reasons there are no observations to support or contradict such conjecture. North of the eastward Circumpolar Current there is evidence of the anticyclonic subtropical gyres which dominate the low- and 31 middle-latitude circulation. The gyre stands out particularly well in the western South Pacific, where the 50-cl/t surface reaches its maximum depth of nearly 2000 m. The trough which marks the axis of the gyre runs along U9°S approximately. Strong southward flow along the eastern edge of the New Zealand Plateau closes the gyre to the west. At shallower depths, the structure of the South Pacific gyre is rather different from that on the 50-cl/t surface. On Reid's (1965) map of the 125-cl/t surface (depth 500-700 m) the gyre en- compasses a much larger area, and its axis lies to the north along 23°S. The western boundary of the gyre is found along the coast of Australia (the East Australian Current), and there is only a slight indication of southerly flow along the east coast of New Zealand. On Reid's 80-cl/t map the axis of the gyre has shifted south to about U0°S. At the depth of this isanosteric surface (1000-1200 m) , New Zealand and the shallow plateau surrounding it form a major obstacle to the flow, and the subtropical gyre begins to divide into two gyres, one in the Tasman Sea and a larger one in the western South Pacific. A poleward boundary current is now clearly present east of New Zealand. Presumably, the ridges north of New Zealand serve to separate the two gyres even more at greater depths , leading to the well-defined closed gyre observed in the western South Pacific on the 50-cl/t surface. In the Indian Ocean, only the western tip of the subtropical anticyclone appears on this map (near 35°S, U0°E) . Associated 32 with the ant i cyclonic gyre is a complex eddy -like feature to the southwest. This appears to be produced by the shear between the intense westward Agulhas Current and the eastward Circumpolar Current. In the Atlantic Ocean too the map does not extend far enough north to show much of the subtropical gyre. However, the poleward western boundary current can be seen along the east coast of South America between 35°S and U2°S. In addition to the ant i cyclonic gyres normally associated with low- and middle -latitude wind-driven circulation in the oceans, there is also a somewhat anomalous gyre south of Australia. North of the elongated trough centered on roughly ^5°S, the 50-cl/t surface shoals, indicating a return flow to the west. Although this feature is evident even in the earliest antarctic sections {e.g.y Deacon, 193T> PL. XIX ) , there is no mention of it in the literature . Systematic hydrographic work in the Australian Bight has only recently been started, and data are still rather sparse. Bye (1968) has discussed some of the characteristics of the west- ward flow, which he called the Flinders Current. According to Bye, the average westward transport of the Flinders Current is 20 - 30 x 10s m3/s. Most of this water comes around the southern tip of Tasmania, Bass Strait being too shallow to permit a sig- nificant flux. The water transported to the west must represent a mixture of water brought from the north by the East Australian Current and antarctic water deflected north by the New Zealand Plateau. 33 The circulation in the Australian Bight is complex. Prom- inent meanders are present at 132°E and 115°E. In Bye's report, these areas- are marked by large-scale eddies. More data will be required to resolve the structure of the circulation in this re- gion. II. B. Salinity Salinity on the 50-cl/t isanosteric surface (Figure h) ranges from 3^.28 X to 3^.62 %, , a difference of 0.31* % . The range of potential temperatures which corresponds to the salinity range for 6Q = 50 cl/t is -1.2°C to 3.1°C At the melting point, water of 6Q = 50 cl/t has the characteristics 9 = -1.86°C, S = 3^.26 %o . The salinity distribution, like the flow, is predominantly zonal, with values increasing from south to north. Perhaps the most striking aspect of the circumpolar salinity distribution is its separation into two distinct bands, a southern one in which the salinity gradient is comparatively steep, and a northern one in which the gradient is much smaller. Roughly speaking, the 31+.52- %o isohaline marks the boundary between the two zones. In the Atlantic and western Indian oceans this isohaline follows the 50 °S parallel. Continuing around Antarctica toward the east, the 3^.52-^,, isohaline shifts toward the south, eventually reaching about 60°S in the eastern South Pacific. The steep salinity gradient is found south of the 3^.52- %, isohaline. The gradient is generally confined to a narrow band approximately 5° of latitude in width. Except in the Ross and 3U Weddell seas, where large areas of uniformly low-salinity water occur, the gradient is contiguous to the antarctic coast. Near the coast, the 50-cl/t surface lies just below the shallow layer of fresh, warm water which forms as the sea ice recedes in summer. Vertical mixing with the navifacial layer provides low-salinity ( ^ 33.9 %> ) water which mixes along the 50-cl/t surface with more saline water from the north, producing the observed gradient. In winter, the character of the salinity gradient is probably not significantly altered, for the 50-cl/t surface intersects the naviface where salinities are less than 3^.0 %, (Gordon and Goldberg, 1970, PL. 3). Relatively fresh water formed at the naviface in summer flows out of the Weddell Sea toward the north and east. North- ward flow occurs east of the South Orkney Islands and along the western side of the South Sandwich Islands. These two outflows combine to form the low-salinity tongue surrounding South Georgia. The cyclonic Weddell Sea gyre is evidenced by the low-salinity tongue projecting toward the east along roughly 62°S. Return flow near the coast is indicated by the westward penetration of more saline water. The anomalously diffuse appearance of the salinity gradient to the east of the Weddell Sea is probably due to the gyral circulation. A second major source of low-salinity water along the antarctic coast is found in the eastern Ross Sea. This water propagates to- ward the northwest, splitting into two branches near T0°S, 180° . One branch continues toward the northwest, extending past the 35 Balleny Islands as far as 57°S, ll+5°E. The other tongue of low- salinity water flows northeast parallel to the Pacific-Antarctic Ridge . The sharp bend in the isohalines at 55°S, lH5°W coincides with the abrupt change in direction of the circumpolar flow noted on the depth map. After crossing over the Pacific-Antarctic Ridge, the Circumpolar Current flows east toward Drake Passage. There is a tendency for some of the water to recirculate into the Ross Sea, although the gyre is not as well developed as in the Weddell Sea. Somewhat higher salinities ( > 3*+. 32 X° , reaching 3^.3^ %» ) than those found in the eastern sector of the Ross Sea are present in the central and western sectors. This is true throughout the water column (Countryman and Gsell, 1966, Figure IT) > not just at 50 cl/t . The higher salinities must be the result of local pro- cesses taking place within the Ross Sea, since the water flowing into the Sea is of lower salinity. The low-salinity tongue extending northward to the east of the Kerguelen Ridge appears to indicate outflow from MacKenzie Bay and the Davis Sea. The abrupt transition from the high-gradient to the low- gradient band is explained by noting that the 3^.52- %> isohaline generally lies just to the south of the steepest slope of the 50-cl/t surface. According to the geostrophic relation, the steep slope marks the high-velocity core of the Antarctic Circum- polar Current. The increased lateral shear and zonal transport associated with the velocity core quickly smear out the lateral 36 salinity gradient, leading to a uniform salinity distribution to the north. Within the uniform circumpolar band, salinities range from 3U.52 %o to about 3^.58 m . The high-salinity water required to maintain these values comes from two principal sources: the western South Atlantic and the western Indian oceans. The first source shows clearly as a tongue of high salinity extending pole- ward along the east coast of South America. Near U0°S, the high- salinity water mixes with lower-salinity water which flows north- ward around the east coast of the Falkland Islands . The resulting mixture (S 'v* 3^.56 %o ) then flows east in the Circumpolar Current. The Indian Ocean source is inferred from (l) the saline water to the south and southeast of Africa, (2) the salinity distribu- tion on Taft's (1963) 60-cl/t map, and (3) the very high salinity values found on the 50-cl/t isanostere in the western Indian Ocean section of Clowes and Deacon (1935). The high-salinity water is evidently derived from North Indian Deep Water which flows south in the western Indian Ocean. Salinity in the Pacific Ocean north of 60°S is exceedingly uniform. Fortunately, all the data plotted in this region were taken by electrical salinometer. Therefore, the small variations that are observed may be used with confidence to trace the circu- lation. The anticyclonic flow around the subtropical gyre is marked by a broad northward displacement of the 3^.56- %■ isohaline. centered on 120°W. Further to the east, salinity increases slight- ly, suggesting slow poleward flow along the west coast of South 37 America. A similar trend (although less well defined because of the lower precision of the data) is evident on Taft's (1963) 60-cl/t map. A distinct tongue of low-salinity water can be seen enter- ing the eastern Tasman Sea from the south and then curving in an anticyclonic sense. This pattern is consistent with the deflec- tion of the Circumpolar Current indicated by the 50-cl/t depth map. The salinity distribution in the Australian Bight is confus- ing. High -salinity water is present at either end of the con- tinent, but the two regions are separated by a well-defined zone of slightly lower salinity between 130°E and 135°E. This is the region where a large cyclonic meander appears on the depth map. There are insufficient data in the northern half of the Bight to be conclusive, but one possible interpretation is that the circu- lation consists of two large eddies, a cyclonic one to the west of Tasmania and an anticyclonic one near 120°E. These eddies would transport low-salinity water north into the center of the Bight. II. C. Dissolved Oxygen The dissolved oxygen distribution on the 50-cl/t surface (Figure 5) is remarkably consistent with the salinity distribution. In particular, the division into two distinct gradient bands observed in the salinity map also occurs in the oxygen distribution. The U.5-ml/l oxypleth will be taken as the approximate boundary be- tween the two bands . 38 South of the U.5-ml/l oxypleth, the dissolved oxygen con- centration increases rapidly. Values exceeding T«5 ml/1 are found nearly everywhere along the coast, except to the west of the Ant- arctic Peninsula, where the dissolved-oxygen concentration does not reach 6.0 ml/1. Maximum values ( > 8.0 ml/l) are found in the southwestern corner of the Ross Sea. Typically, the high- gradient zone is 5° in width, although it is substantially broader east of the Weddell and Ross seas. In the Indian and Pacific oceans, the dissolved-oxygen values decrease monotonically toward the north away from the high-gradient band. Concentrations of less than k.O ml/1 are observed along the northern limit of the maps in both oceans. In the Atlantic Ocean, the concentration of dissolved oxygen reaches a minimum of about k.O ml/1 along U5°S, then increases toward the north. Earlier it was pointed out that the 50-cl/t surface coincides with the vertical minimum of dissolved oxygen over much of the antarctic. Gordon (1967b) also noted the widespread oxygen min- imum, which he identified with the upper Circumpolar Deep Water. Gordon attributed the minimum to depletion by biological activity in low latitudes of the oxygen-rich North Atlantic Deep Water. However, it is clear from Figure 5 that the water flowing south along the east coast of South America is substantially richer in dissolved oxygen than the water in the lateral minimum. The min- 5 This may be due partly to the scarcity of data in this region, especially in the Bellingshausen Sea, where heavy ice is always encountered. 39 imum cannot "be due to the gradual consumption of oxygen in the North Atlantic Deep Water as suggested by Gordon. The source of the oxygen minimum is the oxygen-poor water ( < 3.2 ml/l) found along the west coast of South America (see Barkley, 19&8, Figure 106) . This water flows south and mixes with the circum- polar water in Drake Passage. The low-oxygen water flowing eastward out of Drake Passage stands out as a lateral minimum across the South Atlantic on the 50-cl/t surface and also in the Drake Passage and Greenwich Meridian oxygen sections, Figures 1^, l6. Minimum values in the low-oxygen core in the South At- lantic are: Drake Passage - 3.87 ml/1, U5°W - 3.8U ml/1, Greenwich Meridian - 3.90 ml/1. Off the tip of South Africa additional oxygen-poor water mixes with the eastward-flowing Circumpolar Current. The low- oxygen water which appears as a tongue along 20° - 25°E origin- ates in the North Indian Ocean (Clowes and Deacon, 1935). This water is transported southward along the east coast of Africa, and it feeds the circumpolar flow via the complex eddy observed on the depth map. The effect of this water is offset by mixing with water from the Weddell Sea gyre, which transports high-oxy- gen water eastward along 60°S. As a result, oxygen values in the uniform circumpolar belt do not change significantly south of Africa. High-oxygen water from the Weddell Sea can be seen flowing north into the Scotia Sea along the same paths followed by the low-salinity water. Once again these tongues merge to form a Uo prominent tongue which extends past South Georgia. A high-oxygen tongue along 90°E marks the outflow from Mac- Kenzie Bay and the Davis Sea. MacKenzie Bay is the third largest indentation in the antarctic coast, so it is not surprising to find it acting as a source of oxygen-rich, low-salinity water similar to that originating in the Ross and Weddell seas. In the Ross Sea, the dual tongues of low salinity are matched by a divided tongue of high-oxygen water. High values extend northwest past the Balleny Islands and northeast along the Pacific-Antarctic Ridge. The weak cyclonic gyre north of the Ross Sea is again evident. The anticyclonic gyre in the Pacific may he discerned in the northward penetration of relatively high-oxygen water along 120°W, The southward dip of the oxypleths adjacent to the eastern edge of the New Zealand Plateau represents poleward transport in the western boundary current . The deflection of part of the Circumpolar Current into the Tasman Sea shows up in the high-oxygen tongue reaching as far as U2°S. Oxygen values in the Australian Bight are very uniform. Those in the eastern end off Tasmania tend to be slightly higher than those in the west. However, the differences are too small to allow any conclusions to be drawn regarding the flow. III. THE 30-cl/t SURFACE In the Antarctic Ocean the 30-cl/t isanosteric surface lies with- in the deep salinity -maximum layer nearly everywhere (see Figures 13, IT, 18, 20). The principal exceptions are the areas surround- ing the Ross and Weddell seas, where the salinity maximum is U00- 600 m deeper than the 30-cl/t surface. This is to be expected, since both regions are sources of dense, low-salinity water 6 which alters to a significant degree the distributions around them. As the Ross and Weddell sea waters mix with the saline water from the north, the salinity maximum is eroded and forced to greater depth. The deep salinity maximum is a wide-spread feature of the world ocean. Originating in the North Atlantic, the highly sa- line deep water can be traced through the South Atlantic, across the Indian Ocean south of about 30°S, and into the Pacific Ocean as far as U0°S. The fact that the salinity maximum coincides with the 30-cl/t surface over such a large area of the ocean makes this isanosteric surface of particular interest. At the same time it provides support for the original assumptions upon which this analysis is founded. III. A. Depth The 30-cl/t surface ranges in depth from about 200 m in the Ross and Weddell seas to more than 3500 m in the central South Pacific 6 The high-salinity Ross Bottom Water is much denser than 30 cl/t and does not mix with the deep water. See Figure 33. U2 (Figure 6). In the Atlantic and Indian oceans, the maximum depth of the 30-cl/t surface is slightly over 2500 m. At none of the stations used in this study does the 30-cl/t surface intersect the naviface. It was pointed out earlier that the Deutschtand Weddell Sea stations show navifacial steric anomalies greater than 30 cl/t even in mid-winter. It is probable that this isanosteric surface lies below the naviface throughout the year, although it may be shallower in winter than in summer. In addition to the shallow plateaus surrounding New Zealand and the Falkland Islands, there are several mid-ocean regions where the entire water column has a potential steric anomaly greater than 30 cl/t. In such places the 30-cl/t surface inter- sects the sea floor, which acts as a barrier to flow along the surface . In high latitudes there is a close resemblance between the flow on the 30- and 50-cl/t surfaces. The large-scale deflec- tions of the zonal circumpolar flow associated with topographic features are found in the same places on both surfaces. Weak cyclonic gyres in the Ross and Weddell seas are evident in the topography of the 50-cl/t surfaces and to an even greater degree in the 30-cl/t surface . One high-latitude feature which stands out on the 50-cl/t surface but can not be seen on the 30-cl/t surface is the ant- arctic coastal counter cur rent (East Wind Drift). This fact pro- vides an indication of how shallow the counter current is, since the 30-cl/t surface is only 300 m below the naviface near the coast . U3 A narrow trough in the 30-cl/t surface runs along roughly U3°S south of Australia. While the topography of the isanosteric sur- face indicates anticyclonic relative motion, the absolute sense of flow can not be determined from this evidence alone. This question will he resolved by examining the distributions of properties. The small cyclonic eddy south of Africa appears, like its counterpart on the 50-cl/t surface, to be related to the shear be- tween the westward-flowing Agulhas Current and the eastward Circum- polar Current . In lower latitudes the topographies of the 30- and 50-cl/t surfaces do not match as well as they do in the circumpolar belt. The subtropical anticyclone in the South Pacific, which is so clearly defined on 50 cl/t, is completely absent on 30 cl/t. In view of the great depth of the 30-cl/t surface in this region, it is not surprising that the subtropical gyre is not evident. While the subtropical gyre is not present, there is an indica- tion of a boundary current along the New Zealand Plateau. Stommel and Arons (i960) have predicted from a model of the abyssal circu- lation that a northward-flowing deep boundary current exists in the western South Pacific Ocean. Recent field work (Reid et at.s 1968) has confirmed the existence of the current and suggests a transport of 8 - 12 x 106 m3/s. In the present work, the deep current is reflected not only in the slope of the 30-cl/t surface but also in the distributions of salt and oxygen on that surface and in the distributions in the bottom layer. There is no indication of the subtropical gyre in the Indian Ocean south of 35°S. However, the data are scarce, and the gyre may still persist at this depth. In the Atlantic, the gyre does show up as a narrow trough centered on U0°S. III.B. Salinity The total range of salinity on the 30-cl/t isanosteric surface is from 3^+. 56 %o in the Ross and Weddell seas to 3^.92 % in the western South Atlantic Ocean (Figure 7) • Over most of the map, however, the salinity range is much smaller, about 3^.68 - 3^.80 %o , a difference of only 0.12 X • Potential temperature on this surface decreases from 3.*+°C in the north to -0.8°C near the Antarctic continent. Over most of the antarctic region, the meridional change in potential temperature is about 2°C. At the melting point, water with 6fl = 30 cl/t has the characteristics 9 = -1.8T°C, 5 = 31+.52& . Three water masses with distinctly different salinities converge in the western South Atlantic Ocean. Extremely uniform Circumpolar Deep Water enters through Drake Passage in the west. The salinity of this water varies from 3^.71 %° along the northern side of the Passage to 3*+. 73 %> along the Antarctic Peninsula. After flowing through Drake Passage, the Circumpolar Deep Water continues east until it impinges on the Scotia Ridge, where it is forced to divide. Part of the water flows north out of the Scotia Sea through the passage between the Falkland Islands and U5 South Georgia, while the remainder flows east around the south- ern end of South Georgia. Low-salinity water from the former branch can be seen extending along the continental -shelf of Argentina and along the northern side of the Scotia Ridge. Water of salinity lower than that of Circumpolar Deep Water enters the South Atlantic from the Weddell Sea. The outflow is marked by a tongue of fresh water (S < 3^.68 %, ) centered on the South Sandwich Islands. The effects of this water on the salin- ity distribution can be detected as far north as U7°S. In sharp contrast to the two water types flowing into the South Atlantic from the west and from the south is the highly saline North Atlantic Deep Water. In the tropics, water with salinity exceeding 3*+. 90 %o is found between 2000 and 3000 m across the entire South Atlantic (see Fuglister, I960, p. 6l) . Maximum salinities in the western trough are slightly higher ( *v* 0.02 %o ) than those in the eastern trough. Maximum salinities on both sides of the South Atlantic decrease toward the south, but those on the eastern side decrease more than those on the western side. At 32°S, the maximum in the western high-salinity core is about 0 .0G %o greater than that in the eastern core (Fug- lister, I960, p. 57). Near 35°S the western high-salinity core, which in lower latitudes is pressed against the east coast of South America, separates from the boundary. It mixes with the lower-salinity water from the south and joins the general eastward flow across 1*6 the South Atlantic along roughly UO°S. As it approaches the African coast, part of the deep water turns north in the sub- tropical gyre. The northward flow of this slightly diluted deep water (S = 3^.80 - 3^.8U X ) probably accounts for the greater poleward decrease in salinity in the eastern South Atlantic rela- tive to the western South Atlantic. The deep water which does not turn north in the subtropical gyre continues east past the tip of Africa. This salty water shows up clearly in the northern end of the 20°E vertical section, Figure IT. Near U0°E the 3^.80- %<, is oh aline turns abruptly to the north, indicating that the high -salinity water is flowing along the east coast of Africa. The abyssal model of Stommel and Arons (i960) predicts a northward-flowing deep boundary current in the western South Indian Ocean as well as in the South Pacific. At present the hydrographic coverage of this region is regret- tably sparse, and the prediction of Stommel and Arons remains un- tested. The present map, however, suggests that such a current does exist. Across the Indian sector of the antarctic the 3k-l6-%° iso- haline marks the boundary between the salty deep water originat- ing in the North Atlantic and the less saline Circumpolar Deep Water which flows through Drake Passage. Near 110 °E the salty deep water shifts several degrees to the south before continuing eastward in a broad band across the Indian -Ant arctic Basin. In this region, the vertical salinity maximum associated with the hi 30-cl/t surface develops into a lateral maximum which can be traced around the circumpolar belt as far as Drake Passage (see Figures 18, 20, 13). Maximum values in the high-salinity core are: 132°E - 3^.755 X , 1T0°W - 2h.lk3 % , Drake Passage - 31+.739 %o , which represents a decrease of less than 0.02 %. in about 10,000 km. The fact that a tongue develops in the salinity distribu- tion south of Australia is due to the presence of relatively low- salinity water ( < 3^.7^ %» ) in the South Australian Basin north of 1+Q°S. The Tasman Sea is the only source of such water on this map. Thus, it appears that the deep flow in the South Australian Basin is westward, and the large gyre along U3°S is in fact anti- cyclonic. Transport calculations to be discussed in Section V indicate that the westward flow is extremely weak- in the deep layers . The northward loop in the 3^.1^-Xo isohaline west of New Zealand indicates deflection of some of the circumpolar flow into the Tasman Basin. The rest of the Circumpolar Deep Water flows into the South Pacific and divides into two branches . A portion (8 - 12 x 106 m3/s, Reid et at. , 1968) turns to the north along the eastern edge of the New Zealand Plateau. This tongue of rel- atively high-salinity water marks the beginning of the deep west- ern boundary current which was mentioned earlier. The deep water which flows north in the boundary current fills the entire deep Pacific, there being no sources of deep water in the North Pacific, kQ A northward transport of 10 x 10s m3/s in the boundary current implies a residence time of about 10 years for water in the layer below 2000 m in the Pacific. The bulk of the deep water flows eastward across the Pacific in the Circumpolar Current. This flow is marked by a thin, mean- dering tongue defined by the 3^.7^- %> isohaline. In the eastern South Pacific, water of salinity less than 3^.70 % moves south- ward along the coast of South America. Mixing between this water and Circumpolar Water produces the uniform deep water which enters Drake Passage and flows into the western South Atlantic. Near the antarctic coast, the salinity distribution on 30 cl/t is much like that on 50 cl/t . Low-salinity water originat- ing in the southwestern part of the Weddell Sea moves north along the Antarctic Peninsula and then east parallel to the South Ork- ney Islands. The very distinct tongue along 55° - 60°S shows that this water continues east as far as 20°E before turning south in the cyclonic Weddell Sea gyre. Just to the east of Kerguelen Ridge low-salinity water ex- tends northward from MacKenzie Bay and Davis Sea. In the Ross Sea two tongues of low salinity are again evident, one extending northwestward in the direction of Macquarie Island, the other eastward along the southern flank of the Pacific-Antarctic Ridge. The southward extension of the 3^.7^- %> isohaline along 135°E and the indentation of the 3^.72- %> isohaline along 73°S provide further evidence of a weak cyclonic gyre in the Ross Sea similar h9 to that in the Weddell Sea. This feature stands out in the meridional salinity and dissolved-oxygen sections along 170°W, Figures 20, 21. The high-salinity, low-oxygen water between 500 and 1000 m near 73°S marks the westward return flow into the Ross Sea. (The return flow into the Weddell Sea is evident in the Greenwich Meridian potential-temperature section, Figure 15. Note the relatively warm water near U00 m at 65°S.) III.C. Dissolved Oxygen In the Atlantic and Indian sectors of the antarctic the distri- bution of dissolved oxygen (Figure 8) is distinguished by a lateral minimum ( < U.5 ml/l) . This feature, like its counter- part on the 50-cl/t surface, originates in the eastern South Pacific, where low-oxygen water ( < k.2 ml/l) from the north mixes with higher-oxygen water ( > U.5 ml/l) in the circumpolar zone. The resulting mixture (U.3 ml/l < 0 < U.5 ml/l) flows 2 into the western South Atlantic between oxygen-rich North Atlantic Deep Water and Weddell Sea water, both of which have concentra- tions well over 5-0 ml/l. The flow of Circumpolar Deep Water north into the Argentine Basin is indicated by the northward displace- ment of the oxypleths in this region. As with salinity, the Circumpolar Deep Water and the North Atlantic Deep Water can be distinguished by their oxygen concen- tration across the Atlantic and Indian sectors. The difference in concentration weakens gradually toward the east. South of 50 Australia, the lateral minimum associated with the Circumpolar Deep Water and the weak maximum associated with the North Atlantic Deep Water merge into a broad belt of very uniform oxygen con- centration. Further evidence of westward flow in the South Australian Basin is provided by the oxygen-poor water (0 < U.2 ml/l) north 2 of i+5°S. Low concentrations are also found in the Tasman Basin. The northward extension of the U.5-ml/l oxypleth along l60°E is consistent with the deflection of part of the Circumpolar Current into the Tasman Sea. In the Pacific Ocean, the oxygen concentration decreases monotonically from south to north. Northward flow along the west- ern boundary of the South Pacific is indicated by the prominent tongue of relatively high-oxygen water along l65°W. Dissolved-oxygen concentrations exceeding 5-0 ml/1 occur along the antarctic coast from the Weddell Sea eastward to the Ross Sea. Maximum concentrations ( > 7.0 ml/l) are found in the. southwestern corners of the Ross and Weddell seas and in MacKenzie Bay. High-oxygen tongues extend from the Ross and Weddell seas along the paths followed by low-salinity water originating in these seas . However, the water in MacKenzie Bay has a rather modest influence on the oxygen distribution outside the Bay com- pared with the strong effect on the salinity distribution noted in the previous section. 51 IV. THE BOTTOM WATER Because of the vertical homogeneity of deep antarctic waters, it is not possible to use isanosteric analysis below the 30-cl/t surface. However, it is possible to observe the circulation of abyssal waters in the horizontal distributions of sea-water prop- erties near the ocean floor. The distributions of three such properties — potential temperature, salinity, and dissolved oxygen — have been mapped for this report. Approximately 320 data points were plotted for each map (see Figure 9 for station positions). All sample points are from depths greater than 3500 m, and all but a few are within 300 m of the ocean floor. IV. A. Potential Temperature The most commonly measured and mapped bottom-water characteristic is its temperature, or rather its potential temperature. I have" already noted that maps of bottom potential -temperature, covering part or all of the circumpolar zone, have been prepared by Wust (1933, 1938), Deacon (1937), and Gordon (1966). However, it seemed worthwhile to prepare a new bottom potential -temperature map both for the sake of completeness in this report and because the previous maps either were based on rather poor data grids (Wust, Deacon) or covered only part of the antarctic region. Before examining the map of potential temperature, the ef- fect of geothermal heating on the temperature structure of the bottom water will be considered. Heat generated in the earth's 52 crust is conducted into the water layer above the sea floor. The geothermal flux has been measured at a number of locations in the Atlantic and Pacific oceans. Except over ocean ridges, where it may be somewhat larger, the flux F is uniform and of the order 10~6 cal/cm2s (Bullard, 1963). To find out if this rate of heating will alter to an appreciable degree the thermal structure of near-bottom waters, a simple model of the circula- tion near the sea floor will be constructed. Assume that a thin, vertically homogeneous layer of water ("bottom water") of thickness H flows over the floor of an ocean basin at a uniform speed U. Assume further that the vertical uniformity of the water is due to convective overturning caused by heating from below, and that no mixing occurs between the bottom water and water above it. What temperature change, AT, will occur as a water parcel moves along the basin floor 1 The heat Q added to each water column of unit area as it moves a distance L is Q = — Therefore, the downstream temperature gradient due to geo- thermal heating is (2) AT/L = , QJL = JL pcH pcUH where p = density of sea water c = specific heat. 53 The most difficult to determine of the above parameters is the homogeneous bottom-layer thickness, H (not to be confused with the Ekman boundary layer at the bottom) . Few hydrographic stations with closely-spaced bottles near the bottom are avail- able. Fortunately, some of the stations occupied during Eltanin cruises in the South Pacific Ocean (Cruise 28) and eastern Indian Ocean (Cruises 32 and Ul) do have several bottles at 50 - 100 m intervals near the bottom. These indicate that the bottom few hundred meters of water are extremely uniform. A choice of 100 m, which is typical of the sampling level of the data used in this study, for H would appear to be conservative. The remaining param- eters will be given the following values: F = 10"6 cal/cm2s U = 1 cm/s c = 1 cal/cm3 °C . From (2) the horizontal potential -temperature gradient due to geothermal heating is found to be 0.01°C/1000 km. Observed gra- dients (from Figure 10) are much larger: Atlantic-Antarctic Basin, 0.08°C/1000 km; Southwestern Pacific Basin, 0.lU°C/l000 km; Pacif- ic-Antarctic Basin, 0.2T°C/1000 km. It appears that the geothermal flux does not significantly alter the distribution of potential temperature in the bottom water. In the present map of potential temperature (Figure 10) , as in the earlier ones by Wust and Deacon, the most striking feature 5U is the extremely cold water ( < -0.8°C) in the Weddell Sea. This water, which will be called Weddell Bottom Water, originates along the western shelf of the Sea, then flows east in a narrow tongue along the southern flank of the Scotia Ridge. Some of the Weddell Bottom Water enters the Argentine Basin through the deep South Sandwich Trench. This water continues to flow north along the western boundary of the Basin. According to Wust (1933) , antarctic deep water (Weddell Bottom Water) can be detected in the western trough of the Atlantic as far north as U0°N. Some cold bottom water from the Weddell Sea also enters the Scotia Sea, probably through a passage (sill depth 3000 - 3500 m) near 6l°S, U0°W (Gordon, 1966). A sharp horizontal gradient of potential temperature is found in the Scotia Sea. Reid and Nowlin (1971) suggest that the gradient is caused by the shielding effect of the ridge along 60°W, which allows Weddell Bottom Water to col- lect in the southwestern Scotia Sea. The strong flow of warm deep and bottom water in the Circumpolar Current is deflected northward by the ridge and is confined to the northern half of Drake Passage. The sharp potential-temperature gradient marks the boundary between the warm water from the Pacific Ocean and the cold water from the Weddell Sea. It is evident from the distribution of potential temperature that a substantial part of the bottom water produced in the Weddell Sea continues to flow east along roughly 60°S. Water colder than -0.U°C fills the entire Atlantic -Ant arctic Basin. The temperature 55 increases steadily from -0.8°C in the Weddell Sea to about -0.5°C along the western flank of Kerguelen Ridge. Mixing with warmer deep water accounts for most of the increase in temperature. Although the flow of bottom water in the Atlantic -Ant arctic Basin is primarily zonal, there is some northward flow through gaps in the mid-ocean ridge system. A major break in the ridge occurs near 50°S, 30°E, connecting the Atlantic -Ant arctic and Agulhas basins. Wust (1933, 1938) and Deacon (1937) showed water colder than 0°C penetrating past U0°S, whereas on the present map the 0°C-isotherm reaches only as far as ^7°S. Generally speaking, the bottom water in the Agulhas Basin seems to be somewhat warmer than Wust and Deacon indicated. The distribution of potential temperature in the Agulhas Basin shown in Figure 10 suggests a weaker antarctic influence than would be inferred from the earlier maps . The Kerguelen Ridge, which separates the Atlantic-Antarctic Basin from the Indian -Ant arctic Basin, is very shallow. From kj°S to 62°S the sill depth is less than 2000 m. The only notable pas- sage through which bottom water from the Weddell Sea can flow is at 6U°S, where the sill is at 3000 - 3500 m. East of Kerguelen Ridge, the bottom-water temperature contin- ues to rise slowly as far as 110°E, where it is slightly less than -0.U°C. Beyond that point, the potential temperature drops to nearly -0.6°C. The decrease in potential temperature, along with changes in the other properties to be examined later, is evidence 56 of a second source of bottom water in this region. It will be- come evident that the second source is the Ross Sea. Bottom water from the Indian -Ant arctic Basin flows into the South Australian Basin through the gap in the Indian -Ant arctic Ridge at 120°E. As a result, the potential temperature in the center of the South Australian Basin is slightly lower than it is to the east and west. Some bottom water also flows out of the Indian -Ant arctic Basin through the narrow passage at 59°S, lU6°E. Part of this water moves north into the Tasman Basin while the remainder flows into the Southwestern Pacific Basin over the narrow sill at 59°S, l60°E. Having entered the Southwestern Pacific Basin, the bottom water flows northeast along the New Zealand Plateau in the deep western boundary current and spreads into the central South Pacific. The bottom water south of the Pacific-Antarctic Ridge is sig- nificantly colder than that north of the Ridge. Part of the water in the western end of the Pacific-Antarctic Basin originates in the Ross Sea. The salinity data show that there is also a strong contribution from the Weddell Bottom Water west of the Macquarie Rise. The moderate temperature gradient along the southern flank of the Pacific- Ant arctic Ridge indicates zonal flow from about lUO°W to 100°W. The warmer water observed in the northern Pacific- Antarctic Basin probably comes from the Southwestern Pacific Basin. According to Gordon (19-66), the warmer bottom water crosses the 57 Pacific-Antarctic Ridge through the complex fracture zone located at 130 °W. IV. B. Salinity The map of bottom-water salinity included in this paper (Figure ll) is the first to be constructed for the entire circumpolar region. Previously, the salinity distribution in the Indian and Pacific sectors was regarded as being too homogeneous to provide useful information. However, even the small salinity differences which are observed in this region provide information about the bottom circulation which cannot be derived from the temperature and dissolved oxygen distributions. The bottom salinity has been mapped before in the Atlantic sector of the Antarctic Ocean. A comparison of the old map ( Wiist, 1933) with the present map reveals one significant difference, a difference which led Wiist to draw erroneous conclusions regarding the formation and circulation of bottom water in the Atlantic Ocean. On Wust's map and the present one, low-salinity water ( < 3^.65 %o ) is found along the southwestern margin of the Wed- dell Sea, a region where bottom water is formed. Both maps also show a tongue of low-salinity water extending along the western side of the Argentine Basin. However, whereas the present map shows the salinity increasing steadily from the Weddell Sea source to the Argentine Basin, Wust's does not. On Wust's map a 58 large region of slightly higher salinity separates the low- salinity water in the Weddell Sea from that in the Argentine Basin. This distribution led Wust to conclude that the fresh, cold water in the Argentine Basin derives from a source other than the Weddell Sea. He hypothesized that the bottom water which spreads northward in the western trough of the Atlantic (which he called antarctic deep water A_) forms near the pack- ice boundary in autumn and winter, while the bottom water which forms over the Weddell Sea shelf (which he called antarctic bottom water, A- ) remains trapped in the western Atlantic- Antarctic Basin. I have already stated in Section I.B that the mechanism pro- posed by Wust for the formation of A_-type water is highly im- probable. Oceanographers have for some time accepted the notion that the bottom water which propagates along the western Atlantic trough originates in the Weddell Sea. It is reassuring to find that the salinity distribution is in fact consistent with this view. The salinity distribution helps define the path taken by Weddell Bottom Water as it enters the Argentine Basin. It is clear that the Weddell Bottom Water passes through the South Sandwich Trench, not along the channel east of the Trench (Gordon, 1966). The salinity distribution in the Scotia Sea, like the dis- tribution of potential temperature, reveals an influx of Weddell Bottom Water. The salinity increases rapidly toward the west owing 59 to the influx of more saline Pacific water through. Drake Passage. Examination of several Eltanin stations along the Scotia Ridge between the Falkland Islands and South Georgia showed no bottom water of salinity less than 3^.72 %> passing over the Ridge into the Argentine Basin. The Weddell Bottom Water which enters the Scotia Sea must be thoroughly mixed with Pacific water before it flows into the Argentine Basin. The bottom-water salinity in the Atlantic-Antarctic Basin is highly uniform. Only in the extreme western and eastern ends of the Basin does the salinity differ from 3^- .66 %> . (The apparent uniformity of the bottom water in this basin is partly due to the fact that all the data were taken by chlorinity titration. Salin- ometer measurements will undoubtedly reveal some variations in the salinity distribution.) Low salinities in the southern Agulhas and Crozet basins are indicative of Weddell Bottom Water flowing north through the gaps at 30°E and 60°E. The relatively high salinities found in the northern Agulhas Basin reflect eastward flow around the tip of Africa of water from the South Atlantic. Across the Kerguelen Ridge the salinity of the bottom water increases from 3^.66 %» to 3^.68 %o . Unfortunately, there are no salinometer measurements in the vicinity of the Ridge, and the salinity structure over the Ridge cannot be properly investigated. It is reasonable to assume that in passing over the Ridge the bottom water mixes with the Circumpolar Deep Water, causing an increase in salinity near the bottom. 60 The slight salinity minima in the South Australian and Tasman basins mark the flow of bottom water into these basins. Recall that weak temperature minima were observed in these basins, also. In the South Pacific Ocean, the bottom-water salinity distri- bution is more complex and its interpretation less obvious than in the South Atlantic and Indian sectors. Low-salinity water which crosses the sill near 59°S, l60°E spreads north into the Southwestern Pacific Basin. The tongue of low salinity along the New Zealand Plateau is another manifestation of the deep western boundary current . In the central Southwestern Pacific Basin salinity at first increases slowly, reaching a maximum of about 3^.715 %o near U0°S. Beyond this point, salinity decreases slowly, dropping below 3^.70^ at 30°S. The change in sign of the meridional bottom-water salin- ity gradient is related to changes in the salinity of the deep water. As deep and bottom water flow north after entering the South Pacific, the deep salinity maximum is gradually eroded by mixing. As a result, the salinity of the bottom water increases slightly. North of about U0°S, where the deep salinity maximum disappears in the central South Pacific, salinity decreases monot- onically from the ocean floor to the Intermediate Water. Further vertical mixing tends to decrease the salinity at the bottom. The value of the bottom-water salinity map is well demon- strated in considering the Pacific-Antarctic Basin. The range of salinity in the Basin is very small, and no doubt some of the 6l detail shown in the isohalines is due to observation error. How- ever, I am confident that the pattern shown in the salinity map does reflect real differences in the bottom water. The data grid is denser than in any other part of the antarctic, and the major features have been observed on several different cruises. The essential features of the salinity distribution in the bottom water of the Pacific-Antarctic Basin are: saline water ( > 3^.71 %o ) north of 60°S and in the extreme western end of the Basin x with a band of slightly less saline water ( < 3*+. TO %> ) extending along roughly 65°S. I have already stated on the basis of the potential -tempera- ture distribution that the bottom water found in the northern part of the Pacific-Antarctic Basin probably comes from the South- western Pacific Basin. The salinity distribution is consistent with this view. With respect to salinity, the water in the north- ern Pacific-Antarctic Basin is indistinguishable from that in the central Southwestern Pacific Basin. The small region of relatively high salinity in the western end of the Pacific-Antarctic Basin is the result of outflow from the Ross Sea. For reasons which are not clear, the bottom water formed in the western Ross Sea is much more saline than that in the Weddell Sea. Whereas the cold shelf water of the Weddell Sea has a maximum salinity of less than 3^-70 %> , that of the Ross Sea has a maximum salinity of more than 3^.85 & • Typical station curves for the southwestern reaches of both seas are shown in Fig- ure 33. 62 Finding "... no stream of cold poorly saline bottom water sinking from the Ross Sea," Deacon (193T> p. 115) concluded that the Ross Sea had little or no influence on the bottom water out- side the Sea. Probably Deacon was looking for water like that produced in the Weddell Sea, on the assumption that the similar climatology of the two regions would produce similar water types. Deacon went on to conclude that any bottom water which is formed within the Ross Sea must be trapped by the shallow ridge which runs across the mouth. This thesis has been generally accepted by oceanographers for many years . Recent topographic maps of the Ross Sea {e.g.. Countryman and Gsell, 1966) indicate that the barrier across the outer shelf is broken by several shallow troughs which extend into the interi- or of the Sea. These provide paths for Ross Bottom Water to es- cape. Gordon (1966, 1970) has shown from the distributions of properties outside the Sea that some bottom water does indeed flow out of the Ross Sea. It would appear, however, that he has over- estimated the effect of the Ross Bottom Water, particularly in the western Pacific-Antarctic Basin. He implied that the cold bottom water found there is supplied by the Ross Sea alone. It will be seen that this is clearly not the case. Also, his asser- tion that the bottom water which flows into the South Australian Basin is a mixture of Ross Bottom Water and a "locally-produced water" seems unlikely. If the Ross Sea were contributing large amounts of bottom water, salinities in the eastern Indian-Ant arc- 63 tic Basin would be much higher than those observed there. The conclusion that the bottom water found in the western Pacific-Antarctic Basin is not solely from the Ross Sea is related to the salinity minimum which extends across the Basin at 65°S. Figure 20, which is a salinity section along 170°W, shows high- salinity Ross Bottom Water in the deep basin south of the Pacific- Antarctic Ridge and less saline water ( < 3^.70 %> ) over the crest of the Ridge. Figure 22, which is a salinity section along roughly 65°S from lU0°E to 150°W, demonstrates that the low-salinity water comes from the west. The low-salinity water represents the last traces of Weddell Bottom Water, which flows up the western flank of the Macquarie Rise, along the Pacific-Antarctic Ridge, and then into the western end of the Pacific-Antarctic Basin. Ross Bottom Water can be observed beneath the Weddell Bottom Water at l60° - 170 °E, suggesting that the outflow takes place along the western boundary of the Ross Sea. The fact that the salinity minimum per- sists past 90°W indicates that the flow of Weddell Bottom Water across the Macquarie Rise must be strong relative to the flux of Ross Bottom Water into the Pacific-Antarctic Basin. IV. C. Dissolved Oxygen The map of dissolved-oxygen concentration for the bottom water (Figure 12) is strikingly similar to the potential temperature and salinity maps. The latter two distributions and the circulation patterns they reveal have been discussed in great detail in the 6k preceding sections. To avoid unnecessary repetition, only the most significant aspects of the dissolved-oxygen distribution will be described. The highest concentrations of dissolved oxygen ( > 6.0 ml/l) on this map are found along the western and northern margins of the Weddell Sea. Eastward flow out of the Sea is marked by a tongue of high-oxygen water which is pressed against the southern arc of the Scotia Ridge. Overflow of Weddell Bottom Water into the Scotia Sea is evident in the high oxygen values east of Drake Passage. Oxygen-rich water is also observed in the Argentine Basin. It is clear from the oxygen distribution that the high-oxygen water in the Argentine Basin is derived from the tongue extending north- ward along the South Sandwich Trench. No such water was found to be crossing the ridge separating the Scotia Sea and the Argentine Basin. As Weddell Bottom Water moves east in the Atlantic-Antarctic Basin, the dissolved- oxygen concentration falls from 6.0 to 5.1+ ml/1. In the western Indian-Antarctic Basin the oxygen concentration re- mains at about 5 .k ml/1, but in the eastern end of the Basin the concentration increases to over 5.6 ml/1. This increase, like the small changes in potential temperature and salinity observed in this region, is a result of the influx of Ross Bottom Water from the east. In the Macquarie Rise sector, the oxygen distribution shows bottom water flowing out of the Indian Ocean via three paths: north into the Tasman Basin; across the Macquarie Rise at 59°S and 65 into the Southwestern Pacific Basin; across the Macquarie Rise along 65°S and into the Pacific-Antarctic Basin. The flow into the Southwestern Pacific Basin continues east along the Pacific- Antarctic Ridge and then turns north to feed the deep western boundary current along l60°W. The eastward flow of bottom water across the Macquarie Rise and Pacific-Antarctic Ridge shows up clearly in the oxygen sections along 65°S (Figure 23) and 170°W (Figure 21). A large proportion of this water is Weddell Bottom Water. Ross Bottom Water, with a somewhat higher oxygen concen- tration ( > 5.3 ml/l) , can be seen beneath the Weddell Bottom Water in both sections. In the Pacific-Antarctic Basin, the large region with oxygen concentrations greater than 5-0 ml/l represents both Weddell and Ross bottom waters and their mixtures. The lower oxygen values in the northern Pacific-Antarctic Basin are evidence of the in- flux of bottom water from the Southwestern Pacific Basin and of mixing with oxygen-poor water along the Chile coast. 66 V. FLUX AND VELOCITY STRUCTURE OF THE ANTARCTIC CIRCUMPOLAR CURRENT A shortcoming of the method used in the previous sections is that it provides only a qualitative view of the circulation. In. this section the structure of the circumpolar flow, particularly that of the Antarctic Circumpolar Current, will be examined by more quantitative means. This will be done principally in terms of the relative geostrophic velocity field computed for several meridional hydrographic sections around Antarctica. Deep current- meter measurements taken along 132°E south of Australia will be used to compute the "absolute" net flux in the circumpolar zone, and the result compared with a similar computation of the Drake Passage transport. The mass transport of the Antarctic Circumpolar Current has been estimated by numerous investigators using the dynamic, or geostrophic, method. These calculations have been carried out most frequently for the Drake Passage section, which is consider- ably shorter than any other section across the Circumpolar Current. Clowes (1933) was the first to compute the zonal flux through Drake Passage, which he determined to be 110 x 106 m3/s relative to 3500 db. Later estimates, which range from 90 to 218 x 106 m3/s (excluding the unrealistically low results of Ostapoff, 196l) have been reviewed by Gordon ( 1967a) and will not be discussed here. Sverdrup (19U0), using BANZARE stations along lU5°E, has estimated the net eastward flux from the naviface to the bottom 67 between Australia and Antarctica at 150 x 106 m3/s. Kort (1963) computed the eastward flux in the same sector (l65°E) to be 189 x 106 m3/s from data taken by the Ob' in March 1958. Both Sverdrup and Kort calculated the transport relative to 3000 db at ikk x 106 m3/s. Kort (1963) has also computed the net eastward transport be- tween South Africa and Antarctica. From hydrographic stations occupied by the Ob ' in February -March 1957 he calculated 150 x 106 m3/s relative to 3000 db and 198 * 106 m3/s relative to the bottom7. From Discovery data taken along 20°E in April 1938 Kort computed the transport relative to 3000 db to be 205 x 106 m3/s . While the zonal fluxes computed from the Australian and South African sections are in reasonable agreement, the flux computed for Drake Passage is significantly smaller than the other two sec- tions. Obviously there has been no gain or loss of water in the Atlantic, Pacific, and Indian oceans, since mean sea level has re- mained constant for many years. Simple water budget calculations will show that the mean transports across the Drake Passage, South African, and Australian sections must in fact be very nearly equal. In comparison with the zonal water transport of the Antarctic Circumpolar Current, the average annual difference between evapora- tion and precipitation -plus-runoff makes a small net contribution 7 The substantial increases in transport between 3000 m and the bot- tom which Kort reported seem unrealistically large in view of the very small relative velocities which are observed below 3000 m. 68 to the water budgets of the Atlantic, Indian, and Pacific oceans. For example, an average difference of 1 m/yr between evaporation and precipitation over the entire Pacific Ocean would represent a flux of only 5 x 106 m3/s. The actual average difference, al- though not well known for the oceanic areas, is probably somewhat less than 1 m/yr (see Dietrich, 1963, Figure 75). Aside from the circumpolar zone, the only direct communica- tion among the oceans is through Bering Strait and the East Indian Archipelago. Numerous current measurements have been made in Bering Strait. The mean transport through the Strait is known to be 1.0 - 1.5 x 10 6 m3/s toward the north (see Coachman and Aagaard, 1966, for a summary of Bering Strait transport calculations). The net flux through the East Indian Archipelago has not been deter- mined as precisely as the Bering Strait flux. The only significant passage through the islands is between Timor and Australia. Wyrtki (l96l) has estimated the transport of upper water in the Timor Cur- rent at 1.0 - 1.5 x 106 m3/s toward the west. It is clear from the deep water in the Timor Trench that the deep flow must be in the op- posite direction to the surface current (Sverdrup et al.3 19^2, p. 739). It is unlikely that the net transport through the East Indian Archipelago exceeds 5 x 106 m3/s. The fact that the net transport figures cited earlier do not agree either for a given section or among sections might be inter- preted as evidence of seasonal or longer-term transient motions in the circumpolar circulation. Kort (1963), Deacon (1963), and Gordon ( 1967a) have suggested that transient motions are significant 69 in the Antarctic Circumpolar Current. Recent studies indicate, however, that this is not the case. Boven and Stommel (1968) analyzed sixteen sections along 20°E taken by the Discovery dur- ing the period April 1938 to March 1939. No seasonal variations in the temperature (hence mass) distribution were observed. Reid and Nowlin (l9Tl) have computed the transport relative to the bot- tom through Drake Passage using four hydrographic sections made by different ships between 1930 and 1969 • The range of the trans- ports was 92 - 117 x 106 m3/s; three of the four were from 113 - 117 x 106 m3/s. The author has computed the transport relative to the bottom across two sections south of Australia: 115°E {Discovery , December 1937 - January 1938) and 132°E {Eltanin, December 1969 - January 1970). The net eastward flux through these sections was 1U9 x 10G m3/s and 156 x 106 m3/s, respectively. The results compare well with Sverdrup's (19^0) estimate of 150 x 106 m3/s at lU5°E. It would appear that the net relative transport across any section is reasonably steady, at least with respect to seasonal and longer-period fluctuations. A more likely explanation for the variations in computed transport among the sections is that these calculations determine only part of the flux. The total transport consists of the relative component plus the "barotropic" component (Fofonoff , 1962b) , which corrects for any difference between the true velocity and the assumed velocity at the reference level. For a deep reference level the velocity difference is usually small, TO resulting in relatively minor changes in the computed velocity field. However, a small correction to the velocity field may have an appreciable effect on the transport. A barotropic cor- rection of 1 cm/s between two stations 250 km apart in k km of water represents an increment of 10 x 106 m /s in the transport. In order to determine the absolute transport, current mea- surements must be taken along with the hydrographic stations from which the relative transport is to be computed. Over the past two austral summers two such sections have been made in the antarctic, the first in Drake Passage (January 1969) and the second along 132°E between Australia and Antarctica (December 1969 ~ January 1970). The results of these cruises will be discussed in order to demon- strate the magnitude of the barotropic correction and to estimate the total mean transport. The instruments and techniques used to measure deep currents in the antarctic have been developed at the Scripps Institution of Oceanography. (See Isaacs et al.3 1966, for a detailed de- scription of the current-meter system.) The current meter is tethered to an anchor on the sea floor for a preset period. At the end of the period, the meter automatically breaks its mooring and ascends to the naviface where it is recovered. If the cur- rent measurements are to be used to compute absolute geostrophic transports, as in the present case, the instrument is tethered at or above the level of the deepest bottle on the hydrographic casts on either side of it. In the antarctic experiments, the 71 meters were usually set 300 - 500 m off the bottom, although, two were as much as 900 m off the bottom. Output from the current meter consists of a continuous trace indicating current direction (magnetic) and a record of the number of revolutions made by the Savonius rotor per unit time. The practical threshold of the rotor is 0.5 cm/s . Tests off the Cal- ifornia coast have shown good reproducibility between adjacent meters (Isaacs et al.> 1966). For the present purposes the raw data were reduced to hourly- mean speeds and directions. Semi-diurnal, diurnal, and daily -mean velocity components were derived by computer from the reduced data. In all cases the records were at least one lunar day (25 hours) long. The barotropic velocity component was determined by taking the difference between the measured mean speed normal to the sec- tion and the geostrophically-computed speed at the same depth. For those station pairs with no current meter between them, the barotropic component was determined by linear interpolation be- tween the values on either side. Results of the Drake Passage cruise are discussed in detail by Reid and Nowlin (1971) and will only be summarized here. Five complete current-meter records of 25 - 99 hours' duration were ob- tained. At all current-meter stations the daily mean current was toward the east. Overall mean speeds ranged from about 3 cm/s at the northern end to nearly 9 cm/s at the southern end of the sec- 72 tion. The geostrophic transport relative to the deepest common level of each station pair was found to be 113 x 106 m3/s. The "barotropic transport was 12 U x 106 m3 /s , giving a total eastward flux of 237 x 106 m3 /s . This estimate does not include any cor- rection for the small area outside the sampling grid, i.e., be- tween the continents and the current meters at the ends of the section, or between the sea floor and the reference level. The section along 132°E between Australia and the antarctic continental shelf was made during Ettanin Cruise Ul. Twenty-one deep hydrographic stations were occupied along the section, and five complete current-meter records were obtained. The current- meter results are summarized in Table 2. On the whole, the mean speeds were somewhat smaller than those measured in Drake Passage, as might be expected in the wider section. From the hydrographic data, the geostrophic transport rela- tive to the deepest common level of each station pair was found to be 156 x 106 m3/s. The barotropic component, which was deter- mined in the same manner as for the Drake Passage section, was 77 x 106 m3/s, giving a net transport of 233 x 106 m3/s. Figures 2k and 25 show the relative and absolute velocity and transport fields . It is evident from these calculations that the barotropic component is a significant part of the total transport in the cir- cumpolar belt. In Drake Passage the barotropic component accounted for 52$ of the total, while at 132°E it represented 33$ of the to- tal flux. 73 o en O p ,£3 P ^H O g u O o CM on no o H aJ en % i cu s cu p f (1) o o CO CU * EH o + o o o Eh E-t a o M Q en pq o pq v_^ Cm CQ O PC Eh Ph n eh a O 2 O M En EH EH ft O H CQ Q § CO o ft n K O o o on o o CO o VO O LT\ O on LTN CO w CM VO o ir\ o o O H -3- on l o o ir\ cm co vo o H C— Ov CM CM on ON ON o ON LTN CM CO on -a- CO w ON LTN LTN LTN O O j- on I o o o on cm o on H rH CM o o CM o ON CM VO cn on co w o ON LTN I on CM LTN o o ON CM ON LTN cn on co w O H o o o o O CM Lf\ cn cn t— O H o o O CM ltn cn ltn o -=r o vo ON o o M M r-i cn CM o o cm cn M H on lt\ co LTN on o CM O -3- LTN H cn cn 8 CM -3" CQ W cm -=r o o o o Lf\ CM upv on I LTN O H on CM CM H H H H -4- VO H rH vo CM O CM CM O -H/ VO LTN LTN on on o CM CO w LTN VO o o o o on cm vo en I O LTN O -St -=t t— H H O rH ir\ -p 3 & a o o CD u o «H OJ ^ en xi p c 0) -p o -P nb (l97l)« Investigations of this sort could eventually lead to an understanding of the mechanism(s) involved in the production of bottom water in the antarctic. 86 MAPS (FIGURES 1-12) Figures 2 and 9 show the distributions of stations for the isan- osteric and bottom maps, respectively. The key given below identi- fies the ship symbols. To avoid crowding, only the last two digits of station numbers are shown on some meridional sections. Those stations for which there was no originator station-number in the NODC files are designated by the year the data were taken, enclosed by parenthesis . On the isanosteric maps, areas in which the entire water col- umn is less dense than 50 or 30 cl/t are hatched. On the bottom maps, areas less than 3500 m deep are hatched. Key Eltanin 0 Cruises 28 & 29 (Scorpio Expedition): 0 S Cruise Ul: 0 E Discovery # Ob' A Meteor + Operation Deep Freeze □ Atka: AK Northwind: NW Edisto: ED Staten Island: SI Glacier: GL Westuind: WE Glacier (IWSOE-1968 & 1969) ■ Hakuho-Maru + Thomas Washington A Umitako Maru >{< All others O Anton Bruun: AB Africana II: AF Argo: AR Atlantis: AT CapCtan Canepa: CC Diamantina: D Gascoyne: G Galathea: GA Horizon: HO San Martin: SM Thorshaven: TH Vema: V William Scoresby: WS 87 Figure 1. Major topographic features of the Antarctic Ocean 04»* !<. • ©IBM o *!« j O 30 aIB3l * OTM 17 / rf% „ •'*•* •'«) ./ /M -^..4 ""■•>« .400 v'» — 0^9f§ 014B "" 0W* ^415 0»> ftl» o"« .... ^ °141 64M tin o*» °*"° %» •"' •" .to » ■& » *_ 0» ««I °°* NI OJ^ «' !;»f 'A J* «li»S „**» •"" 0'" 0,,o 0»» »•» *4«« ° Wji SO y «" ^ DN.I»I 5°'l=i$^jC/o "■*> ?;Jo '*' ■* .t. .„rT+ ^ .„ o»« o»«o '8' '.°%«_„ •■!» .i» •» ..«> H «Si 2»0« * O130 _Z30» ■ * «ST ^ ^S» • •'#» ■» V I7*> '3?o tyss o»«o . «*S I4O0 °'» » , .. ••" * .» i~. S I \ ■' « V '»»•• i*» S M M" 0.*"" -r- —* ANTARCTICA ^ j,. STEREOGRAPHIC POLAR PROJECTION SCALE I58«I06(AT 60*S) Figure 2. Stations for the 50- and 30-cl/t maps 89 ANTARCTICA STEREOGRAPHIC polar projection SCALE I 58xl06(AT 60*S) Figure 3. The 50-cl/t surface: depth 90 . * ANTARCTICA * STEREOGRAPHIC POLAR PROJECTION SCALE I58»106IAT 60"S) Figure k. The 50-cl/t surface: salinity 91 } ANTARCTICA ^° STEREOGRAPHIC POLAR PROJECTION SCALE I58»I06(AT 60'S) Figure 5- The 50-cl/t surface: dissolved oxygen ANTARCTICA *>° ^ STEREOGRAPHIC POLAR PROJECTION SCALE I58«I06(AT 60*S) Figure 6. The 30-cl/t surface: depth 93 . * ANTARCTICA *>° STEREOGRAPHIC POLAR PROJECTION SCALE I.58«I06(AT 60*S) Figure f . The 30-cl/t surface: salinity 9h ANTARCTICA *>° + STEREOGRAPHIC POLAR PROJECTION SCALE I 58«106(AT 60'S) Figure 8. The 30-cl/t surface: dissolved oxygen 95 *•'" ^7////// MM Jf Vo 1 ■+ 1 7 "" v v£*>x oHoisn J o»u» ••" 0»M *' •° O** ©«m *W% 0444 .[■/ ^Mw, 04t2T ail &4» *y ^> --W§ X ft MO A 0*50 6?° 1 »». IT / 0*91 + A40 ••■ " %j3 ////// //I' **° 41 Hi 00 \///// .riairci 0.99 f///////M 0«)9 0291 ''iw^mf/ J/ ///A o"*" /////A occ IM Q*/////)\ occ M# X 4194.40 0 6004 0344 ^joo " °° M" *Bx^o*^^i/ " "" /ff/ '///' 0S55 * 0» r« 10! 0*'5 0" MSP^A^M 1 1 **• SW/f/ff/ss^-Lllffff/t °"' ' ^///////X " X- /////i 03ST 0«0 j,, Ji° 030T ? ^r/A)////h\ <»'«*" V. •"*> >- v////j •" 0>" 0« 0" /K-rf 0"i o"9ta w//M/MWh* *"" ., + 0 o'5 .--/ -4- "m* O41* Vy ?a ooY// /// //v] oMi 330 \7// ■ ♦.■'Hij o Y S /////A. ,JT \p///''7/ )/////////// 03 33 M.0^0 V, j .-■ j *12n r/ f\j^y//////k ;,'4S o,^" 0,„ d^uaiw' + 3> 7k *" '"// X \7. °809 r/ 1 X* + + i 90" * W////Mh* o"* v// ///////A ,o x Uu/7///\ to x 'i^isJJ/K 093.^ 0,,, 0334 %/7P * 7 '" //////// ^0. *>, ^ 3. 4i«f {///////j fh^^ ... V//X 04R25 ^ /ff/////////l + + * ////urrr^r •"" + /}/>///////////' + 180° H- 0"+ h2* ";t " w ovZ\ '*^» ////////7/7r>. ao,° °7'' ///////u/////L + + 1 ////A • ////A ;**° *~v>- J 'y/J/j//// . /////r , Y/\ ° // %^ + ■7^J///i ////A x ** +¥V + " J^ ,5M* /7//J1 + ©51! Q0*60l ij o °^'^^^i////\ y ^^^ ih (ff/1 •* 9J OCCHI .... ' ■* '.,. -'^wv; * \J//**%X 4) I99T -V ^9<4 no *"*V///^W^. * >r V////7 x» T- t e o^Cc/////V>, + , ///// v\ 8"o CTO ^W// ^*>»*> VY>X 0€I2 ^< /(*<*// Lr^ •*> iO •isca *»' /////> ~{&t~j v/77/777^ * *"" ,f'* o**'j'A////\ ■A j 131 8B60 inta \y////\ iff • ^*v/// " „» '////////Jh OTTJ^ 4 //////// ^i/f ti«o \y .107 ^Ul//////k x////)i • '•MO ^^f/ ////// AT • ^VV7^>v X * [777777//// X \j///////////l/f f/X DM OH. ••" \ ^W///AV X or? 0 ^\7/7 Y//7rn 1 •)«■• \//////j^ X X / ///////////// Av'vK '"' «.»* ANTARCTICA ^° x ^^ (U o £s a; CO -P CM I »d a; •H o o o CM CM 4 •H -P c 0 -p o •H CO (1) o on CO ON H o h aJ c 0) *-. O co 103 LT\ ld3Q s •H o CO Q o •H •P O 0) CO fl 0) O I % H O CO CO •H >3 •H ■s •H 0) cu u o 10 k /Hid3 a •H §• o o o vo OJ CVJ a ON m ON 3 Pn cd 3 OJ h H 01 in £> CM 0) ^ o o — 05 m CM S s o IO u «M M Cfl •P aJ Q * R e> N • S3 O •H -P m o s > -p • •H ON fl on ■H ON H tH m aJ CM en 43 o fn <« .. a CM w S CM o K O CO < CVJ H Cm 105 8 **/ H ld3Q o o 8 8 8 8 o o o o o — (VI IO <* m 1 1 1 1 1 i I o o o 10 ON Cvl ON VO ON U cu ■§ CU o cu Q CM CM 0) 03 •H 0 Sh O IO S •ri g +i r>» «• fej a (0 o u m «H m 05 c8 « (O • c o •H i*- +3 o cu W >> +3 •H c o> •H H 0) CQ o CM • • IO W o — CM CO rH _ • t-' 00 < l>- H (/) CD g W) fo 106 ld3d CO H 0) •H o CO a) C O •H -P O 0) CO C3 a> o I H O CO CO •H T3 o CM m On •H 107 w/Hid:aa i CO MD ON H U O aJ •H P P-t aJ 3 Q o o o G m O CO •H CM -P H O a> rn W O !>> • <*- P o\ •H MD fl Q\ •H H b crt 3 Is § o i-o o t— On H H O CM 108 Hld3Q O CM CD •H en CD » o a O 10 to «r o5 ■R a) Q • » a » o •H ■P O (O CD *• CO C 0) H o O w CO 3 IO • • 3= o IO o c— t-H r>- IO • » ■p •H fl •H H a5 CO CO o ir\ VD OJ OJ 3 ■H En 110 «*•/ Hld3Q -P Ill 0) ld3Q en 0) o o CO Q ai > •H -P •H CO • o CO ft ■*""«, Vw* •n S !» pm •H o O H o r-i X \i LTN O H •H ,£j CO & 3 ^ 3 -P a* en a; O CD ^ — * M CO o vo •H P 1 a H O cu L^ fH CO «v_-^ O •H «, O CO o 0) bO 0 rH o CO o p a O Sh -h ft CO O CD *H H -p o o •H CD W id a) O 0) T3 CM f-i 3 no cu H H ,£> O LTN CU U •H 113 w / Hld3Q U ■i O O VD •H 43 I ft O O U LTN +i on CO o CD -P no Ti o cp ft > CO 13 H CD o -p • C CO 0) oo C ON O H & 05 3 o o H d c o ISI o [— ir\ on H On C\J CU •H Ilk if\ OJ 00 rH ii ir\ en 0) 03 •3 u 0 c •r* +i f~i Bq s 0 u » -P X •H CJ CT) O CO rH 0) m t> r-i crt 0 3 •H a< J3 CD P< C) .-— -v h en -P 0 CO 0 O VD 0 ■H j- -P ^^ cd Hi £ h 0 ft In m O g P h a p cu C p 0 OJ & s 0 0 • CO H VD aj o\ G H O H h a3 o Q H P O •H H o O H CD > CO o •H 43 — ft CO O O f-t LT\ -p ITS CO O I CM H CO H aJ a* CD £1 ft--- o cn *h o -P C\J W VD O CU I o a; CO > J- •H — • 0) o ■p o ft en p p CD -p c p o -p •H cu J3 & ft o u • -p \£> 01 on o o\ H M £5 ai o t> fn • •H aJ en -P S ■ — , Ct5 ■n H ir\ a 0) H U o «H H o vo X ■p •* c H o CU H ON c o £ CO H g © o5 o » ^ o 0 o< 0 CD H to 03 3 CO O o N s LT\ O VO *-l • • «H 1 H o kJ O \Q -p UA -=r as -rf H n — ' O m CD g ElO •H fe 118 w / Hld3Q O o m O m < 43 o u a C\J CM ON (1) w •H u u s o u en a sr P < CO -— ' on (D •H *■ -P W H 0) c (0 u o CD •H tH -p O ai h +j ^3 pp. 126 REFERENCES (continued) Rossby, C.-G. and collaborators (1937) Isentropic analysis. Bull. Am. Met. Soc, 18: 201-209- Scripps Institution of Oceanography /Woods Hole Oceanographic Institution (1969) Physical and chemical data from the Scorpio Expedition. SIO Ref. 69-15, WHOI Ref. 69-56, 89 pp. Seabrooke, J. D., G. L. Hufford, and R. B. Elder (19T1) Forma- tion of antarctic bottom water in the Weddell Sea. Jour. Geophys. Res., ]6: 216U-2178. Stommel, H. and A. B. Arons (i960) On the abyssal circulation of the world ocean-II. Deep-Sea Res., 6_: 217-233. Sverdrup, H. U. (1931) The origin of the deep water of the Pacific Ocean as indicated by the oceanographic work of the Carnegie. Gerlands Beit. Geophysik, 29_: 95-105- Sverdrup, H. U. (1939) Lateral mixing in the deep water of the South Atlantic Ocean. Jour. Mar. Res., 2_: 195-207- Sverdrup, H. U. (19^0) British, Australian, New Zealand Antarc- tic Research Expedition: Hydrology-discussion. BANZARE Rep., Ser. A, Vol. Ill, Pt. 2, Sect. II, p. 89-126. Sverdrup, H. U., M. W. Johnson, and R. H. Fleming (19^2) The Oceans. Englewood Cliffs: Prentice-Hall, IO87 pp. Taft , B. A. (1963) Distributions of salinity and dissolved oxygen on surfaces of uniform potential specific volume in the South Atlantic, South Pacific, and Indian oceans. Jour. Mar. Res., 21: 129-1U6. Tokyo Univ. Ocean Res. Inst. (1970) Oceanographic data of KH-68-U (Southern Cross Cruise) of the Hakuho Maru, 67 pp. Tolstikov, E. I. (ed) (1966) Atlas Antarktiki, Vol I. Moscow: Acad. Sci. U.S.S.R., XXIII pp. text, 225 pp. figures. Tsuchiya, M. (1968) Upper waters of the intertropical Pacific Ocean. The Johns Hopkins Ocean og. Studies, No. h. Baltimore: The Johns Hopkins Univ. Press, 50 pp. von Arx, W. (1957) An experimental approach to problems in physical oceanography in Physics and Chemistry of the Earth, Vol. 2 (L. H. Ahrens , F. Press, K. Rankama, and S. K. Runcorn, eds). New York: Pergamon Press, pp. 1-29- 127 REFERENCES (continued) Welander, P. (1959) An advective model of the ocean thermocline . Tellus, 11: 309-318. Wooster, W. S. and B. A. Taft (1958) On the reliability of field measurements of temperature and salinity in the ocean. Jour. Mar. Res., IT: 552-566. Wooster, W. S. and G. H. Volkmann ( I960 ) Indications of deep Pacific circulation from the distribution of properties at five kilometers. Jour. Geophys . Res., 65_: 1239-12^9. Wright, R. (1969) Deep water movement in the western Atlantic as determined by use of a box model. Deep-Sea Res., Fuglister Volume: 1+33-^6. Wust , G. (1933) Das Bodenwasser und die Gliederung der atlantischen Tiefsee. Wiss. Ergebn. dt . atlant. Exped. Meteor j 6_: 1-107. (Eng. trans., U.S. Naval Oceanog. Office TRANS- 3^0, 1967.) Wust, G. (1936) Atlas zur Schichtung und Zirkulation des atlan- tischen ozeans. Wiss. Ergebn. dt . atlant. Exped. Meteor, 6, Atlas, Teil A & B, 92 Beil. Wust , G. (1938) Bodentemperatur und Bodenstrom in der atlantischen, indischen, und pazifischen Tiefsee. Gerlands Beit. Geophysik, 5kj 1-8. Wyrtki , K. (1961) Physical oceanography of the Southeast Asian waters (NAGA Report, Vol. 2). Scripps Institution of Oceanography, 195 pp. VITA Jeffrey Edwin Callahan was "born in Cambridge, Massachusetts on 2U September 19^3 • He graduated from the United States Naval Academy in June 1965s receiving a Bachelor of Science degree and a commission in the United States Navy. After two years' duty aboard the USS Willis A. Lee (DL-U) in the Atlantic Fleet, he entered the graduate program in oceanography at The Johns Hopkins University. He received a Master of Arts degree from the Univer- sity in 1969. The author is married to the former Evelyn Green of Bethesda, Maryland. Thesis •C1889 1 •> Thesis .C1889 Callahan The structure and circulation of deep and bottom waters in the Antarctic Ocean. 129027 IhesC1889 The structure and circulation of deep an 3 2768 002 08452 7 DUDLEY KNOX LIBRARY