The Two Faces of the North Atlantic Oscillation By John M. Wallace 1. Introduction The atmosphere-ocean system has exhibited some rather pronounced trends during the past few decades. Winters have been getting milder, particularly over northern Europe, Russia, and central and western Canada. Global warming skeptics take delight in pointing out that the strongest warming trends have been observed over the coldest regions of the hemisphere. One notable exception to the trend toward global warming has been Labrador and west Greenland, where winters have been growing more severe. Barometric pressure has been dropping over the Arctic and the westerlies at subpolar latitudes (~55°N) have intensified. Winter storms over the North Atlantic have been getting more intense and wave crests have been getting higher. Winter precipitation over Scotland and Norway has increased. Snow accumulations at higher elevations in Norway have been so heavy that a number of the glaciers have begun to advance, even in the face of the trend toward milder temperatures. Meanwhile, winter rainfall over southern Europe has been decreasing, with major droughts during the early 1990's and again last year. Winter temperatures in the lower stratosphere 20 kilometers above the earth’s surface have been cooling over the Arctic polar cap region, with temperature drops of 5 C or more, and there has been appreciable thinning of the stratospheric ozone layer poleward of 40°N. In this lecture I will argue that all the trends noted in the previous paragraph are all linked to the same phenomenon and that equally pronounced changes over and around the Antarctic are linked to its Southern Hemisphere analogue. 2. Background The challenge of understanding climatic change is promoting a remarkable synthesis of the earth sciences. A generation ago, when I entered the field, geology, meteorology and oceanography were largely separate disciplines, and each of them touched on only limited aspects of the climate problem. The study of climates of the past was largely the province of geologists who, in collaboration with paleo- botanists and zoologists were assembling an increasingly detailed picture of the record of past climate variations based on proxy evidence. Some of them carefully scrutinized their proxy records of past climate with hopes of finding evidence of orbital and other periodicities might lend some insight into the underlying causes of climate variability, but beyond that, most of the scientists working in this field weren't in a position to do much more than gather evidence and speculate. They lacked the tools to perform a global synthesis of the proxy records and to investigate the dynamical mechanisms that link climate variations in widely separated parts of the world. The meteorologists were developing the tools needed for investigating climate variability. They were already having some success in using numerical models based on fundamental physical principles to predict time evolution of the global wind field, but at that time the skill of weather forecasts was limited to a few days. By performing extended runs with the same numerical models, researchers were able to realistically simulate many of the features of the present climate, a notable example being the transport of water vapor in the global hydrologic cycle, as documented by the late Professor Jose Peixoto of this university. Yet despite the promising developments in climate modeling, I think it is fair to say that most meteorologists at that time tended to view climate as unchanging, and climatology as a stagnant field. It was only a few of our peers who recognized the great potential of numerical models for the study of climate variability. A generation ago much of the research on the oceans was focused on regional phenomena. The more global thinkers in the field were still struggling with fundamental issues, like the factors that govern the strength of the western boundary currents, and the rate of recirculation of water between the surface layer of the ocean and the deep abysses below. Ocean general circulation models were just in their infancy. Yet already, developments within the field of oceanography were helping to set the stage for a renaissance in climate research. Marine sediment cores were providing a wealth of detailed information on past climate variability. Leading chemical oceanographers of that time were beginning to consider the carbon cycle and other biogeochemical cycles from a global, interdisciplinary perspective. And pioneering efforts in coupled atmosphere-ocean modeling were already on the drawing boards. 3. The El Nino Southern Oscillation phenomenon 1966, the year I entered the field also marks the publication date of the first of a series of highly influential papers of Jacob Bjerknes [1] on what has come to be referred to as the El Niño-Southern Oscillation (ENSO) phenomenon. ENSO provides what is perhaps the clearest example of how the atmosphere and oceans, as a coupled system, are capable of supporting a mode of space / time variability altogether different than the modes of variability of the individual components of the system. ENSO involves an atmosphere-ocean feedback loop, as shown in Fig. 1. During the onset of a warm 'El Niño' event, weakening of the equatorial Pacific tradewinds leads to a deepening of the thermocline on the eastern side of the basin and decreased upwelling, both of which enable equatorial sea surface temperatures to warm toward the ambient temperature of the tropical oceans. Raising the temperature of the surface waters imparts additional buoyancy to the overlying air, enabling it to support deep cumulus convection. The increased convective rainfall along the equator favors a further weakening of the trades, which allows sea surface temperature to warm further--- a positive feedback. The redistribution of the tropical rainfall also affects the upper air circulation at higher latitudes, producing a complex array of temperature and rainfall anomalies at the earth's surface, as illustrated in Fig. 2, in which the color shading shows the region of enhanced rainfall, as indicated by infrared satellite imagery, and the contours show the anomalies in the temperature of the troposphere, as inferred from satellite microwave imagery. During winters of 'El Niño years' when the surface waters of the equatorial Pacific are abnormally warm, temperatures over Alaska and western Canada tend to be warm as well, as shown in Fig. 3 and regions of the Americas at subtropical latitudes tend to be wetter than normal. El Niño years are also usually marked by drought over Northeast Brazil, Central America, and parts of Africa, and a suppression of hurricanes over the Caribbean Sea. The equatorial Pacific sea surface temperature anomalies that occur in association with El Niño events typically persist for about a year: much longer than the characteristic 'memory time' of either the winds or the equatorial ocean currents alone. The longer time scale is thus attributable to the coupling between the atmosphere and the ocean. By virtue of this extended memory, the coupled system is far more predictable than the atmosphere alone. This extended predictability is being exploited to make forecasts of El Ninorelated climate anomalies out to several seasons in advance, using state-of-the-art coupled atmosphere-ocean models. The principal beneficiaries of these forecasts are residents of tropical countries, many of whom suffer drought during El Niño years, and residents of the countries of North and South America. The effects of ENSO on European climate are more subtle, but the forecasts are nonetheless of some value. The ENSO paradigm has been instrumental in demonstrating the existence of a certain degree of order in the year-to-year variations in temperature and rainfall in the tropics and over some parts of the extratropics. It has also revealed important linkages between the physical variables and the biosphere. For example, the deepening of the thermocline in the eastern Pacific during the El Niño years results in a decreased availability of nutrients to the euphotic zone, to the detriment of many species of marine life. But the increased rainfall along the coastal drainage basins of Ecuador during these intervals provides increased nutrients for shrimp farms the coastal estuaries and habitat for mosquitoes. For the same reasons, sea birds languish while land birds thrive on the Galapagos Islands during El Niño years. These ENSO-related variations in the biosphere are of considerable interest in their own right, and they provide valuable clues as to how the physical variables in the climate system (temperature, ocean currents, winds, etc.) influence marine and terrestrial ecosystems, not only on the year-to-year time scale, but potentially on much longer time scales as well. Around 1976-77 the ENSO cycle underwent a discernible shift. Within the 30-year interval prior to that time, cold phases tended to be slightly more intense and more prevalent than the warm phases, whereas in the years since then, the opposite has been the case. This 1976-77 'regime shift' is more strongly apparent in North Pacific climatic time series and in various indicators of the state of the marine ecosystem [2]. There are indications of an analogous shift, but in the opposite sense around 1946, and perhaps another just within the past 2-3 years, coinciding with the extended cold phase of the ENSO cycle that has persisted from 1998 until the last month or two. This interdecadalscale ENSO-related variability is also sometimes referred to as the Pacific Decadal Oscillation. Whether it is a distinctive phenomenon in its own right, or merely a peculiar feature of the ENSO cycle remains to be seen, but it's generated considerable interest in the fisheries community because it appears to be strongly correlated with the decade to decades ups and downs in salmon recruitment along the coasts of Alaska, British Columbia and Washington. Research on ENSO is continuing. Efforts are underway to improve the skill of ENSO prediction, to extend the record of the ENSO cycle farther back into the past using proxy records, to quantify the impacts of ENSO on physical climate variables, on marine and terrestrial ecosystems, and on human activities, to understand the nature of ENSO-like variability on decadal time scale, and to predict whether and how global warming might impact ENSO. But even as research on ENSO continues, another phenomenon that may ultimately prove to be equally important from a global perspective, and perhaps even more important from a European perspective, has been attracting increasing attention-- a phenomenon variously known (in different segments of the research community) as the North Atlantic Oscillation, the Arctic Oscillation, and the Northern Hemisphere annular mode. 4. Early history of the North Atlantic Oscillation The observed tendency for winter temperatures over Greenland and Scandinavia to vary out of phase (i.e., for one to be above normal while the other is below normal) can be traced back 200 years or more [4]. The realization that this distinctive pattern of temperature anomalies is linked to an equally distinctive north-south see-saw of sea-level pressure anomalies between Iceland and a broad east-west belt centered on Portugal, dates back roughly 100 years. The French meteorologist Tesserenc de Bort, the German meteorologists Julius Hann and Felix Exner, and the British applied mathematician / meteorologist Sir Gilbert Walker all contributed to its discovery. Figure 4 shows Exner's rendition of the pressure and temperature patterns, as represented in a paper published in 1913 [5]. The contours represent correlations with a time series of winter sea-level pressure over the Arctic. During winter months when the pressure over the Arctic is above normal, the pressure over Portugal and much of temperate latitudes tends to be below normal, as indicated by the belt of negative correlations centered over Portugal. In a similar manner, when the pressure over the Arctic is below normal, the pressure over Portugal tends to be above normal. Since the winds tend to follow the pressure contours with low pressure to the left, it follows that below normal pressures over the Arctic favor abnormally strong westerlies over the North Atlantic from Nova Scotia toward Scotland and Scandinavia and vice versa, as illustrated in Figure 5. Hence, the pressure pattern defined by Exner's correlations can equally well be characterized as a modulation in the strength of the westerlies across the North Atlantic shipping lanes and, more generally, at subpolar latitudes. The corresponding surface air temperature correlation pattern indicates that high pressure over the Arctic is conducive to above normal temperatures in Greenland and above normal temperatures over northern Europe and vice versa. Exner reasoned that the temperature and pressure patterns are related by way of the winds: e.g., abnormally strong westerlies that prevail when Arctic pressure is below normal carry an abundance of warm air off the Atlantic into northern Europe causing temperatures to be above normal there. The influence of the winds upon the temperatures is clearly apparent in Fig. 5: when the westerlies are strong the western limit of subfreezing temperatures, indicated by the edge of the bluish shading, lies near the border between Germany and Poland, whereas when the westerlies are weak subfreezing conditions extend all the way into France. It was Sir Gilbert Walker who coined the term 'North Atlantic Oscillation' in a paper published in 1924 [6]. Walker also coined the term, 'Southern Oscillation', which is the SO in ENSO, and the less enduring term 'North Pacific Oscillation' describing a suite of statistical relationships in the Pacific sector. By attaching names to suites of statistical relationships, Walker elevated them to the status of phenomena, at least in the minds of those who read his papers and took them seriously. In the case of the Southern Oscillation, his intuition has proven correct: the statistical relationships embodied in that name have proven to be the signature of the response of the global atmosphere to a distinctive mode of atmosphere-ocean interaction in the equatorial Pacific. But it was not until the work of Bjerknes and his successors, nearly half a century after Walker coined that term, that proof was forthcoming. Prior to that time, believing in the Southern Oscillation required something of a leap of faith. In an exchange of correspondence some years later, Walker was challenged to defend his belief. When one of his contemporaries gently chided him about the lack of a plausible mechanism that would explain the global pattern of correlations that he identified with the Southern Oscillation, he replied: "I think the relationships of world weather are so complex that our only chance of explaining them is to accumulate the facts empirically... there is the strong presumption that when we have the data of the pressure and the temperature at [altitudes of] 10 and 20 km, we shall find a number of new relationships that are of vital importance" [8]. Walker was absolutely right: upper air data proved to be instrumental in understanding the ENSO phenomenon. His intuition regarding the North Atlantic Oscillation has also withstood the test of time. Beginning in the late 1930's Carl Gustaf Rossby (regarded by many as the father of dynamical meteorology) and collaborators at Massachusetts Institute of Technology promoted the concept of an 'index cycle', which they hoped would prove to be a useful tool in forecasting weather on the 1-2 week time scale [9]. This notion was based, not on the prior work cited in previous paragraphs (which they were apparently unaware of at the time) but upon the fact that (1) observations that the strength of the prevailing westerlies in middle latitudes fluctuates from one week to the next and (2) their simplified analysis of the governing equations suggested that these relatively slow fluctuations should be reflected in the configuration of planetary waves, which determines the longitudes where temperatures are above or below the longitudinally averaged values. For example, one week, when the westerlies are abnormally strong, Europe might lie within a cold trough in the planetary wave pattern, whereas a week later, when the westerlies are weaker, it would lie within a warm ridge. The 'zonal index' was originally defined as a measure of the strength of the westerlies that encircle the North Pole at latitudes ~45°N. By the late 1940's Rossby had lost interest in the index cycle, but in 1951 two younger members of the same department published papers that seemed to imply that this concept still held promise. In one of the papers [10] Jerome Namias proposed a redefinition of the index cycle: instead of monitoring the strength of the westerlies as in the earlier papers, Namias argued that it would be more informative to characterize the hemispheric circulation in terms of the mean latitude of the westerlies. Based on his experience as a synoptic meteorologist, he reasoned that when the westerlies are poleward of their mean position, cold, Arctic air masses tend to be confined within the polar cap region, and temperatures in midlatitudes tend to be above normal. Conversely, when the westerlies are displaced equatorward of their mean position, incursions of cold air masses into middle latitudes are more severe and more frequent. As a measure of the mean latitude of the westerlies Namias used an index based on the difference between the strength of the westerlies along 55 N and 35 N. He used the term 'high index' as characterizing the state of the hemispheric circulation when the westerlies are north of their mean position, and 'low index' as characterizing the state when they are south of their mean position. The other paper [11] was one of the first publications of a new professor by the name of Edward Lorenz: a statistical analysis of the correlations between the pressures and winds at various latitudes. Lorenz's study was reminiscent of Exner's study nearly 40 years earlier, but for one important distinction: the pressure data were averaged around latitude circles before the correlations were calculated. But despite this difference in the methodology, Lorenz's results were remarkably similar to those of Exner, which he wasn't aware of. In particular, Lorenz confirmed that sea level pressure over the polar cap region tends to be negatively correlated with pressure along 45 N (the latitude of Lisbon). And consistent with the more subjective impression of Namias, he reported that this favored pattern in his statistical analysis is well represented by an index consisting of the strength of the winds averaged around 55 N. In retrospect, it is clear that the North Atlantic Oscillation (NAO) and the index cycle are different ways of describing the elephant pictured in Fig. 6. Namias's 'high index' and 'low index' states have subsequently been used by many authors used as a way of characterizing the contrasting states of the NAO, where 'high index' refers to below normal sea-level pressure over Iceland and enhanced westerlies across the North Atlantic, along 55 N. Consistent with Namias's findings, the high index state is indeed associated with a easterly wind anomalies along 35 N. In the Atlantic sector to the southwest of Portugal it is manifested as a strengthening of the tradewinds [Fig. 7]. 5. The Renaissance Later in the 1950's the index cycle fell out of favor for lack of evidence of simultaneous week-to-week variations in weather in different sectors of the Northern Hemisphere. When I was in graduate school at MIT during the 1960's it was barely mentioned. The North Atlantic Oscillation fared somewhat better. In a paper published in 1978, Harry van Loon and his student Jeffery Rogers drew attention to its strong influence on winter temperatures over North America, as shown in Fig. 8 [12]. Another of van Loon's former students, James Hurrell documented its strong influence on precipitation over Europe (Fig. 9) [13]. Over Portugal and much of southern Europe most of the wintertime rainfall occurs when the NAO is in its low index state. Over Scotland and Scandinavia the situation is just the opposite: it is the high index state, characterized by strong westerlies off the North Atlantic that accounts for most of the rainfall. Over the years, several of us have also shown, by means of a statistical analysis technique known as empirical orthogonal function (EOF) analysis, that the elephant pictured in Fig. 6 is also the most pronounced mode of variability of the Northern Hemisphere sea-level pressure field; i.e., it accounts for more of the month- to-month variability of sea-level pressure averaged over the hemisphere than any other single pattern, including the one associated with ENSO [14]. Starting with Hurrell's paper in 1995 there has been a meteoric rise in the number of research papers devoted to the NAO and its various pseudonyms, which I believe is attributable to recent discoveries relating to: (1) the strong resemblance between the NAO and the principal mode of variability in the Southern Hemisphere (2) the NAO-like signature in observed climate trends over the past few decades and, (3) dynamical linkages between the NAO and the wintertime polar night jet in the stratosphere. These discoveries occurred almost simultaneously, so I've chosen to present them in thematic, rather than chronological order. 5.1 The NAO's austral twin Meteorologists have long been aware that the circulation of the Southern Hemisphere exhibits much greater circular symmetry about the pole than that of the Northern Hemisphere. The greater symmetry is attributable to preponderance of oceans equatorward of Antarctica, which prevents the buildup of cold, continental air masses analogous to those that cover Siberia and Canada during the winter months. Like the mean circulation, the principal mode of month-to-month variability tends to be circularly symmetric or annular. Moreover its structure exhibits a remarkable resemblance to the 'index cycle' described by Namias and Lorenz, with a 'seesaw' in sea-level pressure between the polar cap region and temperate latitudes [Fig. 10] and a seesaw in the strength of the westerlies between latitude belts centered near 35 and 55 S [15]. A strong resemblance is also apparent in the temperature field. Hence, the NAO has an austral twin. And because this twin exists in an environment devoid of strong land-sea heating contrasts, it is living proof that the existence of the NAO transcends the peculiarities of the Northern Hemisphere land-sea distribution. In trying to describe the significance of this realization, I have likened have it to discovering that the newspaper reporter whom one has known for years by the name 'Clark Kent', is really Superman. In several of our papers, my former graduate student, David Thompson and I have gone so far as to argue that the name 'North Atlantic Oscillation' is a misnomer; i.e., that from the point of view of the governing dynamics, a more global name like 'Arctic Oscillation (AO)' or 'Northern Hemisphere annular mode' would be more appropriate for describing Exner and Walker's pattern than the name 'North Atlantic Oscillation (NAO)', which singles out a particular longitudinal sector of the hemisphere [16]. Whether the baby should be renamed at this point is debatable, but is becoming increasingly clear that this pattern plays a role in climate variability throughout most of the Northern Hemisphere, as envisioned in Namias's articulation of the 'index cycle' paradigm. This pattern, whatever we choose to call it, does, indeed, orchestrate the occurrence of cold air outbreaks and winter storms not only in Europe, but also throughout most of the Northern Hemisphere [7]. These influences weren't apparent to meteorologists back in the 1950's because they didn't have access to the analysis tools or the extended historical records of past data that we have the benefit of today. 5.2 AO/NAO-related climate trends James Hurrell’s 1995 paper was the first the first to point out that the pronounced trends in wintertime climate since the late bear a striking resemblance to the patterns that we see in the month-to-month variations associated with the AO/NAO. For example, the pattern of falling pressures over the Arctic and rising over temperate latitudes resembles the AO/NAO signature [Fig. 11]. The same is true of surface air temperature. Changing wind patterns account for roughly half the observed temperature rises over the interior of Eurasia. [17]. The AO/NAO signature is also clearly apparent in the trends in winter precipitation. If these trends are, indeed, related to systematic changes in the atmospheric circulation that resemble the month-to-month changes associated with the AO/NAO, then instead of thinking of them as being comprised of many separate elements as described above, it is more meaningful to characterize them as manifestations of a shift in this pattern toward its high index polarity, as shown in Fig. 12. Putting all the eggs in one basket, so to speak, greatly simplifies the problem of attribution, since in this case we are justified in focusing on the single question of what is causing the shift in this pattern. But before we consider this question, let's look at the other factors that are contributing to the resurgence of interest in this phenomenon. 5.3 The stratospheric connection Another important development of the past decade is the growing appreciation of the strong linkage between the AO/NAO and the so called 'polar night jet ', a major feature of the circulation of the lower stratosphere. The polar night jet encircles the Arctic polar cap region during wintertime, the season of perpetual darkness, and it blows from west to east. It's a direct result of the strong heating contrast between the dark polar cap region and the sunlit lower latitudes. Under certain conditions, natural or human induced changes in the strength of the polar night jet can impact climate at the earth's surface. Let me cite a few examples. Meteorologists have noticed that major volcanic eruptions like that of Mt. Pinatubo back in 1991 tend to be followed by relatively warm winters over much of middle latitudes. The temperature pattern and the associated wind and pressure patterns bear a strong resemblance to the distinctive signature of the AO in its 'high index' polarity: sea level pressure is below normal over the Arctic, the westerlies across the North Atlantic and, more generally along 55°N are abnormally strong, etc [18]. The apparent source of these anomalies is a global cloud sulfate aerosols formed by the injection of sulfur dioxide into the lower stratosphere. The aerosol particles absorb incoming solar radiation, warming the lower stratosphere. The notable exception is the polar night region where there is no sunlight to absorb. The differential heating between the sunlit and dark regions of the global stratosphere strengthens the equator-to-pole temperature gradient which, in turn, strengthens the polar night jet. The strengthening the polar night jet, in turn, initiates a kind of 'chain reaction' that leads to a strengthening the westerlies along 55°N all the way down to the earth's surface, as shown schematically in Fig. 13 [19]. The AO is also impacted by midwinter breakdowns of the polar night jet. In contrast to the southern hemisphere, polar night jet, which is almost perfectly circular and remains strong throughout the winter, the northern hemisphere jet is continually being perturbed and distorted by upward propagating planetary-waves, driven by the thermal contrast between the cold continents and the warm oceans and by the deflection of the westerlies by the Rockies and the Himalayas. From time to time the planetary-waves over high latitudes become so strong that the ring of cold air enclosed by the jet splits jet into pieces that drift away from the pole and mix with the surrounding air. These breakdowns of the jet are first evident in the upper stratosphere near the 50 km level, and they propagate downward to the base of the stratosphere around 10 km above the earth's surface over the course of a week or two. Once a major breakdown occurs, it takes radiative cooling over the polar cap regions a month or two to restore the jet to its former intensity. During this extended recovery period, the AO has been shown to exhibit a distinct preference for its 'low-index' polarity with above normal sea-level pressure over the Arctic [20]. These conditions have been shown to favor wintry weather conditions in northern Europe, Asia and North America [21]. The polar night jet and the AO both appear to be impacted by variations in the winds that circle the globe in the equatorial stratosphere, which fluctuate back and forth between easterly and westerly with remarkably regularity. The mean period of these equatorial wind fluctuations is a little longer than two years: hence meteorologists refer to them as the quasi-biennial oscillation or QBO [22]. It has been observed that breakdowns of the stratospheric polar night jet tend to occur more frequently when the quasi-biennial oscillation is in its easterly phase. The remarkable regularity of the QBO confers a certain degree of predictability out to a year in advance that appears to be roughly comparable, in terms of skill and economic value, with the predictability associated with El Niño, as documented in Fig. 13. The QBO could conceivably prove to be the more important predictor for European climate, which is more strongly impacted by the AO/NAO than by El Niño. Coincident with the trend in the AO toward its high index polarity over the past 30 years, the winter stratosphere has been getting colder and the polar night jet has been getting stronger and less easily pushed around by perturbations in the planetary waves. (In this sense it's becoming more like its counterpart in the southern hemisphere). Midwinter breakdowns haven't been occurring as frequently and the jets in both northern and southern hemispheres have been persisting later into spring [23]. Increasingly frequently, winter temperatures in the lower stratosphere over the northern hemisphere polar cap region have been dropping below –80 C, the threshold for the formation of polar stratospheric clouds, which serve as a medium for some of the chemical reactions that lead to the catalytic destruction of ozone by freon and other industrial chlorofluorocarbons. Quite likely in response to this trend, there have been indications, within the past few years, of an incipient Northern Hemisphere 'ozone hole' . The cooling of the polar cap and the strengthening of the polar night jet at stratospheric levels over the past few decades is consistent with the observed trend toward the 'high index' polarity of the AO (or NAO) down at the earth's surface, and it could even conceivably be responsible for it . A recent series of numerical experiments by Drew Shindell and colleagues at NASA’s Goddard Institute for Space Studies [24] suggest that human activities might be responsible for the trends at both levels. In their model, as in the real atmosphere, increasing concentrations of greenhouse gas molecules warm the lower atmosphere by trapping outgoing infrared radiation, but they cool the stratosphere by enabling it to radiate energy to space more effectively. As greenhouse gas concentrations increase in their model simulation, the resulting cooling at stratospheric levels tends to be particularly strong during wintertime over the polar cap region. The trend toward colder winters over the polar cap is accompanied by a strengthening of the polar night jet, which favors the 'high index' polarity of the AO (or NAO) down at the earth's surface, with mild wet winters over northern Europe and dry conditions over Portugal. These experiments are suggestive, but not definitive at this point. They have yet to be verified with other models and the mechanisms responsible for the various linkages have yet to be fully elucidated. Plausible ideas have been put forward to explain the linkages, but at this point, not all scientists are agreed that human activity is responsible for the observed trend in the AO/NAO or even that the the trend is due to processes operating in the stratosphere. For example, James Hurrell and Martin Hoerling have suggested, based on their own simulations with a climate model, that the observed trends in both the AO/NAO and in winter temperature and winds in the lower stratosphere could be occurring in response to subtle shifts in sea surface temperature and rainfall patterns in the tropics [25]. 6. Impacts of variations in the annular modes in the Arctic and Antarctic The AO/NAO and its austral twin are playing a role in shaping current research on the polar regions on paleoclimate, and on terrestrial and marine ecosystems. Here are a few illustrations. The AO exerts a strong influence upon the pattern of surface winds over the Arctic and the winds, in turn, control the movement of the sea ice. The prevailing winds drag the sea ice around the pole in a clockwise sense. When the AO is in its high index state (as it was during much of the 1990's) this clockwise gyre weakens, allowing a larger fraction of the ice to exit the Arctic Ocean via the Fram Strait, rather than recirculating in the gyre, as shown in Fig. 14. Under the same 'high index' conditions, the wintertime winds over the coast of Russia have a stronger offshore component, dragging the thick, multi-year ice away from the coast, promoting the formation of much thinner 'first year' ice which melts during summer. Both effects, which are observable on a year-by-year basis, have evidently contributed to the observed decline in ice thickness over the Arctic, and to the increased incidence of open water during summer during the 1990's relative to the previous decade. Figure 15 shows that the area along the Russian coast that has been experiencing a trend toward open water during summer during the past 20-years is virtually identical to the area of reduced sea ice typically observed in summers following winters in which Meanwhile, Antarctic climate has exhibited some strong trends that appear to be related to the AO/NAO's austral twin, the 'Antarctic Oscillation'. In contrast to the situation over most of the globe, surface air temperatures of over the Antarctic continent have not risen significantly over the past few decades and Antarctic pack ice has not been retreating. Yet over the Antarctic Peninsula, temperatures have been rising, ice caps have been melting, and penguin colonies have been relocating farther south. David Thompson and Susan Solomon have shown that the sharply contrasting temperature trends over the continent and the peninsula fit the pattern of the month-to-month variations associated with the Antarctic Oscillation which, like its counterpart in the northern hemisphere, has been residing in its 'high index' state a larger fraction of the time during recent decades. The high index state favors stronger westerlies around the periphery of Antarctica, sweeping relatively mild, marine air over the peninsula, while making it more difficult for marine air masses to penetrate into the interior of the continent [27]. 7. The annular modes and paleoclimate The AO/NAO and its austral twin are evidently responsible for some of the large regional climate trends of the late 20th century. Could they have played an analogous role in the larger climate swings that have occurred on time scales of centuries, millennia and longer? There are several intriguing possibilities. The correspondence the coldest interval of the Little Ice Age and the "Maunder Minimum" of the sunspot cycle during the 17th Century has fueled speculation that the relative absence of solar ultraviolet emissions during this extended period might have been responsible for the harsh climate during this period. This idea was initially greeted with skepticism. Meteorologists readily acknowledge that the sun's plays a critical role in many processes in the stratosphere, but it constitutes a negligible fraction of the solar radiation penetrates into the lower atmosphere. Hence a decrease in solar ultraviolet emission, in and of itself, would not make climate noticeably cooler at the earth's surface. Drew Shindell and collaborators [28] argue that a weakening of the stratospheric circulation that occurred in response to decreased ultraviolet radiation could have affected climate at the earth's surface through its dynamical influence on the AO. Tree ring data and other proxy evidence examined in their paper suggests that climatic conditions observed during the Maunder Minimum, with harsh winters over Northern Europe bear a marked similarity to the 'low index' polarity of the AO/NAO, consistent with the results of their numerical experiments. The AO and its austral twin has also been mentioned in connection with the much larger temperature swings that have occurred in time scales of tens of thousands of years in response to subtle cyclic variations in the earth's orbit. Alex Hall [29] has found that the spatial patterns of sea-level pressure differences between glacial and interglacial climates, as simulated by a climate model, bear a strong resemblance to these 'annular' or 'ring like' modes. On even longer time scales, it's tempting to speculate that the absence of cold winter weather at subarctic latitudes during the Cretaceous period 70 million years ago might have been the reflection of an ultra high index state of the AO. But lest I overstate the case, it should be emphasized that not all the large climatic swings of the distant (or not so distant) past bear the earmarks of the AO/NAO. For example, the rapid temperature rise over Northern Europe during decade of the 1920's, which is sometimes attributed to the trend toward the 'high index' polarity of the NAO, was accompanied by warming of comparable magnitude over west Greenland. If the NAO had been primarily responsible for the warming over Europe during this period, West Greenland should have cooled, rather than warmed. 8. Impacts on ecosystems The impacts of the AO/NAO on marine and terrestrial ecosystems are extensive. Impacted species include shrimp, squid, lobsters, cod, salmon, seals, caribou, musk ox, moose lynx, seabirds, songbirds, wheat, grapes, olives, a variety of flowering plants and trees [30]. Interactions between climate scientists and ecologists is benefiting both disciplines. Ecologists can test various hypotheses correlating their year by year time series of species abundance and growth rates, nesting dates, flowering dates, etc. and various sites with indices of the AO/NAO. Climate scientists, in turn can use ecological data such as tree ring thicknesses and blooming dates as proxies for climate data and thereby extend their historical indices farther backward into the past. 9. Concluding remarks Some instructive parallels and contrasts can be drawn between the AO/NAO and ENSO. ENSO is inherently an equatorial phenomenon, but it impacts winter climate over large areas of the extratropics. In contrast, the AO/NAO and its austral twin exhibit largest amplitudes over high latitudes, but their signatures extend all the way across the tropics, into the subtropics of the opposing hemisphere. ENSO and the the AO/NAO both vary on time scales much longer than the chaotic waves and vortices that dominate the circulation of the lower atmosphere and consequently are to some degree predictable well beyond the 1-2 week limit of conventional weather forecasts. Although some important details remain unresolved, I think it's fair to say that the basic mechanism of the ENSO cycle is relatively well understood. In contrast, fundamental issues concerning the nature of the AO/NAO are still under debate: e.g., is it inherently sectoral or annular? what gives it its distinctive shape? Is it more than just a statistical entity? In the case of ENSO, the alternation between warm and cold states (El Niño vs. La Niña) has been of primary interest, whereas the AO/NAO has been of primary interest in relation to global climate change. It remains to be seen whether the pronounced trend toward the high index state, which began in the late 1960's is indicative of a secular trend that will continue into the future, or whether it is merely one rather dramatic realization of random variations from decade to decade. Only time will tell. Acknowledgements To be added References 1. J. Bjerknes, Tellus, 18, 820 (1966). J. Bjerknes, Mon. Wea. Rev., 97, 163 (1969). J. Bjerknes. J. Phys. Oceanogr., 2, 212 (1972) 2. N. J. Mantua, S. R. Hare, Y. Zhang, J. M. Wallace, and R. C. Francis, Bull. Amer. Meteorol. Soc., 78, 1069 (1997) 4. H. van Loon, J. C. Rogers, Mon. Wea. Rev., 106, 296 (1978) 5. F. M. Exner, Sitzungberichte der Mathematische-Naturwissenschaftlichen Klasse der Akad. Wissenschaften, 122 Abteilung 2a, Zweiter Halbband, 6(10)1165 (1913) 6. G. T. Walker, Mem. Ind. Meteorol. Dept., 25, 275 (1924) 7. D. W. J. Thompson, J. M. Wallace, Science, 293, 85 (2001) 8. R. B. Montgomery, Mon. Wea. Rev., 39 (Suppl.) 1 (1940) 9. C. -G. Rossby, J. Mar. Res., 3, 38 (1939) 10. J. Namias, J. Meteorol., 3, 130 (1950) 11. E. N. Lorenz, J. Meteorol., 8, 52 (1951) 12 D. W. J. Thompson, J. M. Wallace, J. Climate, 13, 1000 (2000) in tradewind figure 13. later van Loon paper 14. J. E. Kutzbach, Mon. Wea. Rev., 98, 708 (1970). K. E. Trenberth, Mon. Wea. Rev., 109, 1169 (1981). J. M. Wallace and D. S. Gutzler, Mon. Wea. Rev., 109, 784, TW98 15. D. Gong and S. Wang, Geophys. Res. Lett., 26, 459 16. D. W. J Thompson,., J. M. Wallace, Geophys.Res, Lett., 25, 1297 (1998). R. A. Kerr, Science, 284, 241 (1999). J. M. Wallace, Quart. J. R. Meteorol. Soc., 126, 791 (2000) 17. D. W. J. Thompson, J. M. Wallace, and G. C. Hegerl, J. Climate, 13, 1018 (2000) M. P. Baldwin and T. J. Dunkerton, Science, 294, 591 (2001). 18. P. M. Kelly, P.D. Jones, J. Pengqun, Int. J. Climatol. 16. 537 (1966) 19. K. Kodera., J. Geophys. Res., 99, 1273, (1994). A. Robock and J. Mao, Geophys. Res. Lett., 12, 2405 (1992). 20. M. P. Baldwin and T. J. Dunkerton, Science, 294, 591 (2001). 21. Thompson D. W. J., M. Baldwin, J. M. Wallace, J. Climate (2001) 22. R. J. Reed, W.J. Campbell, L.A. Rasmussen and R.G. Rogers, J. Geophys. Res., 66, 813 (1966) 23. D. L.. Hartmann., J. M. Wallace, V. Limpasuvan, D. W. J. Thompson, and J. R. Holton, Proc. Nat. Acad. Sci., 97, 1412 (2000, J. M. Wallace and D. W. J. Thompson, Physics Today, 2, 28 (2002) 24 D. T. Shindell, R. Miller, G. Schmidt, L. Pandolfo, Nature, 399, 452 (1999) 25 M P. Hoerling and J. W. Hurrell, Science, 292, 90 (2001) 26. I. G. Rigor, J. M. Wallace and R. L. Colony, J. Climate, in press (2002) from caption of Fig. 27. D. W. J., Thompson and S. Solomon, Science, in press (2002) 28. D. T. Shindell, G.A. Schmidt, M.E. Mann, D. Rind and A. Waple, Science, 294 (2001) 29. A. Hall, Amer. Geophys. Union Fall Mtg., A22B-10 (2001) 30. M. C. Forchammer and E. Post, Trends in Ecology and Evolution 15, 286. Figure captions Fig. 1 ENSO Feedback loop 15 Fig. 2 Hemispheric impacts of the 1982-83 and 1997-98 El Nino events. 16 Fig. 3 ENSO SAT pattern 17 Fig. 4 Exner's SLP map 27 Fig. 5 European maps from Science paper 28 Fig. 6 Elephant 43 Fig. 7 tradewinds 37 Fig. 8 AO and SAT N. American perspective 40 Fig. 9 AO and precip: European perspective 41 Fig. 10 Austral twin SLP 45 Fig. 11 SLP; trend and AO from Science article 50 Fig. 12 AO time series (updated) 53 Fig. 13 AO schematic 59 Fig. 14. Sea ice motion ams 68 Fig. 15 summer sea ice concentation ams 69