Jump to content
Snow?
Local
Radar
Cold?

Leaderboard

Popular Content

Showing content with the highest reputation since 19/04/23 in Articles

  1. There's been a lot of focus on 850mb temperatures over the last few days and will they/won't they support snowfall, so just wanted to create a post for any of the less experienced members who are probably left scratching their heads about the conflicting views amongst different members and want to know what to believe. Let's start with why everyone looks for the -6c 850mb (or hPa) line and it's a bit of a history lesson I'm afraid. I believe in terms of these (and other) forums, the magical figure of -6c was really coined as the "snow line" back in the bad old days of the late 90's and early 00's over on the old BBC Snowwatch forums, when things were so desperate that we were crying out for any sort of 2 day northerly toppler just to look forward to (which is why I do find it amusing that some members can't see the potential in the upcoming period, I think unfortunately December 2010 distorted the expectations of some). Anyway, because back in those days we were primarily looking for Northerly topplers for any sort of cold, given an almost complete lack of Easterlies between 1996-2005, the 850mb temperature needed to be significantly below 0c for the airmass to be conducive for snow falling (this is a bit of an oversimplification, but let's keep it light!). Why? Well in the lower part of the troposphere we generally expect that as we travel from the clouds to the surface of the earth, the temperature increases. However we also have to remember that cold air falls, and so what we expect is the colder air from the 850mb level to fall down closer to the surface of the earth over time. The rate of this temperature increase (or "thermal gradient") varies hugely due to an overwhelming number of factors, however if we consider this in a simplified form again, the answer lies, for us at least, in the oceans. Because from a North or North-Westerly airflow the air is travelling a long distance over a generally warm Atlantic ocean, there is more likelihood of warmer pockets of air close to the surface of any airflow from the N/NW cancelling out some of the colder air dropping from aloft than there would be from say the East, where the air is travelling a far shorter distance over the North Sea and so is less likely to include these warmer pockets of air. These warmer pockets of air can affect many of the parameters that we won't go into in this post, but the two that are worth calling out as they are mentioned a lot are the 2m temperature and the 2m dew point. So if we try and summarise the above two paragraphs, in a N/NWly airflow the air is travelling over warmer oceans and so warmer pockets of air can be expected to be found closer to the surface. To override these warmer pockets of air we need the cold air aloft (at the 850mb level) which falls down towards the surface of the earth to be cold enough to override these warmer pockets of air and make the air from cloud to surface cold enough throughout to support all of our lovely snowflakes falling from the clouds remaining as snowflakes rather than falling through a warmer pocket (or layer) of air and melting into horrible rain, which from experience is much more difficult to make snowballs from. I think the primary reason we chose the -6c line was that back in those bad old days where Wetterzentrale was the choice of most members to consume their daily dose of GFS from, the dashed isotherm line which indicated the 850mb temperatures were spaced out at approximately every 5c (although for some reason I seem to remember Wettzentrale often showing -6c instead of -5c as one of the dashed isotherm lines, correct me if I'm wrong if anyone can remember), and so we would often look for the -5/-6c isotherm on the 850mb temperature charts as our guaranteed snow line, as it allowed for enough headroom for some warmer pockets of air at the surface to be overridden by the falling colder air from above. At this time I think it's fair to say that none of our collective weather knowledge was anything like it was today, and so looking for something simple like a single "snow line" helped us in our search for snow. So, on to the next part, what exactly is the 850mb (or hPa - they both essentially mean the same thing) temperature chart. Again I will keep this as simple as I can, so apologies to anyone offended by the oversimplified statement I may be about to make. Well to answer the first part, it's exactly what it says - it is the temperature of the air at the point in the atmosphere where the air pressure is equal to 850 millibars (mb) or hectopascals (hPa). But the exact height above the surface of the earth of where pressure is equal to 850mb can change, and that is what all of the pretty colours on the "height" charts we see are. So we've been talking about the 850mb level of the atmosphere, let's take a jump to the other common charts we see - the 500mb charts, such as the one below: The colours here represent how high above the earth the 500mb pressure level is. The more towards blue/purple the colour is, the lower the height (or closer to the surface of the earth) the 500mb pressure level is, and vice versa for the greens/yellows/oranges. This also has a knock on effect on how high/low the 850mb pressure level below it is. So what we are saying in the context of our upcoming cold spell is that under that huge low pressure system, the height above the surface of the earth of both the 500mb and the 850mb pressure level is lower. What this means is that the colder air aloft, at the 850mb level, does not have as far to fall down towards the surface, and therefore more colder air is able to come down and help override any warmer pockets/layers of air towards the surface. This means that we don't necessarily need the 850mb temperature to be below -6c to support snow falling to the surface. We also then need to bear in mind that as that low pressure falls further South, and we change our feed of air from being from the North/North-West across the Atlantic ocean to instead coming from the East from the near continent, fewer of these warmer pockets of air are likely to exist. Just to give you an example of this from the most recent 6z run, here are two 850mb temperature charts, the first one at +138 hours, and the other at +168 hours: It would stand to reason that if it was as simple as colder 850mb temperature = colder surface temperature, then we would expect to see a lower 2m dew point temperature in the corresponding +168 chart right? Well see for yourself: You can clearly see that despite the 850mb temperature being 1-2c warmer at +138 than at +168, the dew point temperature is 1-2c lower at +138c. You can again see the reason for this by looking at the 500mb height charts: We can see here that the lower heights at +138 aid the cooler 2m dew point temperatures. So in summary for what is a very long post, you do not require -6c or colder 850mb temperatures, even in a North-Westerly airflow sometimes, in order to guarantee all of the relevant parameters for snow being the right side of marginal, it is significantly more complex than that, and that is why the upcoming period could well deliver a lot of surprises - the word could being crucial in there!
    4 points
  2. Hi. I just wanted to create a post from an educational perspective and also be a point of reference regarding the importance of the GSDM and the AAM on the medium and long term. Hopefully, anyone who questions this should change their mind after reading the below. The reason it is worth doing this is because there is no better, clearer, example of the importance and connections between those upstream, Pacific developments and the downstream influence. I'll attempt to highlight and stitch this together using some of the usual plots and charts which tell the story so very well indeed and it is this which, hopefully, helps people to at least better understand the influences of the GSDM and what to look for moving forward. OVERVIEW: Firstly, we have just endured a very wet December, this comes as no surprise given what happened to the AAM in late November. Interestingly, seasonal models all pointed towards December being a +ve NAO month, in keeping with the long-term teleconnections with regard El Nino and a +IOD in early winter. The late Nov and early Dec cold spell can be linked back to the behaviour of the GSDM too, while what followed through much of December certainly can and what is happening now and looking ahead, most certainly is. Despite a weakened stratospheric polar vortex, it is the troposphere that is 'leading the dance' rather than the stratosphere for the most part even though it is lending a hand. The late Nov/early Dec blocked pattern has links to the temporary rise in the AAM in mid-November, with the usual lag in place. We can see this first evolution in the below plots... The second and more obvious evolution was then the more pronounced fall in AAM through late Nov as the MJO returned to the W Hem and westerly inertia was removed from the GSDM budget and, as usual, AAM fell along with a -FT (Frictional Torque) and -MT (Mountain Torque) event. The result was, as ever, propagation through the extra-tropics and then into the mid-latitudes of more westerly momentum, propped up by easterly momentum at circa 30-40N - Remember the inflated Azores high in the run-up to Christmas? - We can see how the +AAM anomalies, in particular, were distributed through the atmosphere within the usual window of approximately 14 days. No surprise we then saw an enhanced period of +AO and +NAO conditions just before mid-December onwards. Lastly, comes the 'big event' within the last week or so in terms of the pronounced rise in AAM. Once again, in complete contrast to what happened through mid-November, with a lag of course, we can see how the marked rise in AAM has allowed the propagation through the sub-tropics and into the mid-latitudes of more easterly inertia (-AAM) very well indeed, this is particularly obvious on the relative AAM transports plot, as highlighted below. Once again, no surprises at all and of which has been documented by the usual few on the group we are now seeing a complete reversal of that +AO and +NAO period with an array of mid-latitude and, eventually, high latitude blocking patterns. We can also see this visually on the zonal wind plots at 100hPa (top of the trop/bottom of the strat) too. While like in many areas of meteorology, there is never often a usual "a+b=c evolution", but this is about as close to that as you can get. We have certainly had help from the stratosphere here mind because a robust stratospheric polar vortex that is downwelling westerly winds into the troposphere can often be the dominant player, overriding what has occurred over the last month. There is likely to be some 'help' here from the El Nino and eQBO combination this winter as the sPV continues to remain disorganised and far weaker than it can be at this point in the winter with far less influence on the troposphere too. When it comes to the AAM it is crucial to be able to accept and acknowledge when there are other overriding influences and counterbalances. Tamara et al often talk about not taking the MJO at 'face value', it is similar with the AAM, but when you combine the usefulness of the GSDM and all it incorporates then, as I have perfectly examined here, nobody can ever say the likes of the GSDM and the AAM is "flawed or useless" when it comes to pre-empting NWP, at times, and also gauging how the broader patterns may evolve and shift. As I mentioned at the start of the post, I wanted to put this one together because there is no better example of the usefulness of the GSDM than what has occurred over the last month or so. The evolution and 'story' of the atmosphere have been well played out in these plots. For those who want some winter weather after the last few weeks of very wet conditions then 'this is your time'. We approach mid-winter with such solid footing for cold synoptics that it should be a pleasure to watch the more unusual patterns being modelled in NWP and, interestingly, this continues to link in well with the majority if not all of the seasonal models with regards to how Jan and Feb and perhaps even Mar should progress. Winter is about to start... With regards, Matt. Original post: https://community.netweather.tv/topic/99706-model-output-discussion-into-2024/?do=findComment&comment=4994387
    2 points
  3. The Cloud Works Clouds are simply condensed water droplets known as water vapour. These clouds can achieve many different things; they can give warnings, they can produce preciptation or merely tell how stable/unstable the atmosphere is. Clouds can come in many forms from the fluffy fair-weather Cumulus to the thick billowing frontal cloud of the Nimbostratus. Fair-weather Cumulus are caused when the sun causes heating of the ground and this heat known as convection causes clouds to grow on the boundary between cool air and warm air and sometimes these fluffy fair-weather Cumulus can grow into Cumulonimbus causing heavy midday showers or evening thunderstorms. Frontal cloud is the same process where warmer air bumps into cooler air or vica versa, causing moisture to form where the two air masses meet. This is also controlled by evaporating water from the sea/ocean before hand but only when a temperature gradient meet does condensation become very active and these two split areas of heat are called fronts. The Cloud Types Cirrus Cirrus are wispy clouds. These clouds are made of ice crystals and sit at the top of the layer of the atmosphere called the Troposphere where all the weather takes place. Cirrus is usually a sign that a frontal system containing strong winds is on the way or they can be found in the anvil of a Cumulonimbus. These clouds are known also as 'Mares' Tales' and are found around 4-10 miles up from sea level. Cirrostratus Cirrostratus are sheets of stratus like formation high in the sky. These are not particularly a sign of bad weather to follow. The sheet is made mostly of ice and is often seen in winter following on from periods of snow where the sun is visible through the cloud. Sometimes these can produce isolated snow flurries in winter. Again these clouds are 4-10 miles up from sea level. Cirrocumulus Cirrocumulus are uniformed sheets of cumulo-form cloud usually made of ice. These clouds indicate higher level instability particularly in summer and can be a sign of a coming thunderstorm. In winter these clouds often precede snow showers and can again give their own flurries if it is cold enough. These clouds are usually made of ice. The clouds are also known as 'Mackeral Sky' due to the fish-scale apperance they take on. These clouds are usually found around 4-5 miles up from sea level. Altocumulus Altocumulus are patches of cloud usually found in a uniformed formation and these clouds are usually made of ice or water droplets. These are very good sign of mid-high level instability and can sometimes tell of a thunderstorm or heavy shower on its way. In the winter these clouds can tell of storms and snow following but these clouds do not give any precipitation. These cloud are usually found 2-6 miles above sea level. Altostratus A sign that a front is on the way, these clouds can tell whether there will be periods of rain or snow and are often the cloud that precedes Nimbostratus on a warm front. These clouds are thick and they take on a sheet formation of water droplets or ice crystals. These clouds can even produce hours of snow themselves under the right conditions. These cloud again are found about 2-6 miles above sea level Cumulonimbus These clouds can tower as high as mount everest and are famous for their thunderstorms. A Cumulonimbus consists of a Towering Cumulus and an anvil of cirrus on the top. Although the bases of these clouds are less than 2.5 miles above sea level they can extend about 10 miles into the higher reaches of the Troposhere. The clouds are made of ice crystals and water droplets and give heavy rainfall and hail and sometimes snow associated with a cold front. Cumulus These clouds are usually fluffy fair-weather clouds built up by convection but enough sunlight can transform them into Towering Cumulus which can give torrential afternoon downpours. These clouds are made of water droplets though a Towering Cumulus can have ice crystals as the main feature at the top. These are mainly a shower cloud and can give showers of rain or snow. These clouds again form less than 2 miles above sea level. Nimbostratus These are deep layers of stratus usually associated with warm or occluded fronts. These clouds give precipitation of rain or snow and sometimes even hail and have been known to produce lightning. These clouds are slow moving and have a shallow temperature gradient associated with the warm front and usually indicate stable air is on the way. These clouds are made of water droplets and are found around about 2 miles above dea level. Stratus These clouds are low level, uniformed sheets of thick grey cloud. These clouds often appear with weak fronts and are usually a sign of stable air. Such clouds produce drizzle or light rain and when they hit ground level they are known as fog. These clouds are particularly associated with High Pressure areas of the Atlantic in the winter and summer where the land is cool and the sea is warm. These clouds are made from water droplets and form below 2 miles above sea level. Other Clouds They are the cloud types that are the mechanics of our atmosphere but there are other clouds including Mammatus which are found under Cumulonimbus anvils during a death of a severe thunderstorm. Also Pileus clouds which are a good indication of a strong shear these are wispy clouds that over ride Towering Cumulus clouds. Of course the noctilucent clouds that produce eerie shape and colours that are crafted by split colours from the sun or other sources of light.
    1 point
  4. Hi all I promised to give this guide, so here it is. Its not complete and it is for those who are new to trying to work out this very difficult question. What to look for to get snow at sea level. (1) in showers (2) frontal weather (1) In showers 1) Dry bulb temperature below 5C, often 3C is a better mark 2) Dewpoint at or below zero 3) wet bulb temperature, if you have a weather station, no more than about 2C 4) 1000-500mb thickness (DAM) less than 522dm, lower if you are on the coast, but as high as 540dm it is possible in a heavy shower, but unlikely. 5) 850mb temperature of -7C or below, -5C it can occur but not often. 6) on the 850mb chart if the value shown on the contour line is below about 1290dm or 1300dm, then there is a high chance that ppn will be of snow. (The Met Office use 1293dm for a 50% and 1281dm as a 90% chance of snow) 7) zero degree isotherm or freezing level of 1,000ft or less to give a 50% or higher chance of snow at sea level. At 2,000 ft above sea level the chance is reduced considerably. (2) Frontal Weather Two types of front (1) warm front (2) cold front In (1) warmer air is flowing over the top of cold air, and in the case of (2) cold air is undercutting the warm air ahead of it. (1) is, I suppose the classic heavy snow situation which with strong winds can give blizzard conditions, even on relatively low ground, assuming all the factors are in its favour. This assumes they are, namely that very cold air lies near the surface but is not being moved away by an approaching frontal system. This stagnates and eventually retreats away again. In this instance then the above requirements need to be met along with (a) If prolonged and fairly heavy ppn occurs then what we call the wet bulb temperature will start to lower and this can enhance the probability of snow falling. In this instance if the wet bulb temperature is at 3C or below then the ppn can turn to snow from rain. Also, and this applies to showery conditions also, if the wet bulb freezing level(not easy to find on any chart!) is 2,000 ft or below, then ppn can readily turn to snow. In light ppn then, often, regardless of any of the above factors being favourable drizzle or light rain will fall not snow. (2) In this instance then if the air is not all that mild in the so called warm air and the cold air is very cold, with near negative values close behind the front, then even moderate ppn will readily turn to snow as the cold air undercuts the mild air. Use the values above for a guide. This has only touched the surface of trying to forecast will it snow or not but I hope gives, the less experienced, a guide of what to look for. On the topic of wet bulb temperature. I have still not worked out how to give Paul sufficient data that they can set up an algorithm to produce charts of wet bulb potential temperature and also wet bulb freezing level charts. If anyone knows how to do this please pm me or Paul as it would be another big step for Net Weather to be able to produce such charts from the GFS run. I hope you have lots of opportunity this winter to try this guide out. Footnote by Blessed Weather: This article was written by John Holmes, retired Met Office professional, in 2015.
    1 point
  5. Dewpoint: The strict definition for this is:-The temperature to which the air must be cooled in order that it is saturated with respect to a water surface and for this to occur at its existing pressure and humidity content. Dewpoint may be measured indirectly from wet- bulb and dry-bulb temperatures with the aid of humidity tables. More often now it is produced by some kind of calculator or computer programme taking these basics into consideration. Dew forms on clear nights when there is little or no wind at the surface. It can occur on summer nights but is most prevalent during the long nights from autumn through to spring. Dew forms on surfaces whose temperature falls to below the dewpoint of the air in contact with it. From what has already been said it should be clear that the dewpoint is always below the surface temperature. As in many things to do with Meteorology there is a caveat. Remember that the definition is 'become saturated with respect to a water surface'. There are a very small number of occasions when, with temperatures falling, the relationship becomes to that with respect to ice. In these few instances, briefly, the dewpoint may be a fraction higher than the air temperature. Dewpoint is really a measure of how much water vapour there is at any one time. Obviously the higher the air temperature then the higher the dewpoint can be. Compare the temperature in winter say of 3C with that on a hot summer day, maybe 25C. Obviously much more water vapour is possible with higher temperatures. This is partially responsible for the intensity of thunderstorms in summer compared to those of winter, and to the Tropics having much more intense downpours than in Temperate latitudes. The frost point is that temperature at which the air is saturated with respect to an ice surface. Dew point is closely associated with humidity. Thus in warm frontal zones the humidity is high. The arrival of a cold air mass will usually bring a sharp drop in the dewpoint. Dewpoint is an important tool for forecasters to use when forecasting many weather phenomena, be it thunderstorms, human comfort levels, or the likelihood of snow or frost. by John Holmes
    1 point
  6. Polar Lows Before you read this article I would suggest you read my article on Air Masses: Air Masses And the guide by @Thundery wintry showers on Uk winter setups Winter Setups Having read all that lot then I hope this fairly short article will help an understanding of Polar Lows. How they form and an idea of how to predict their track Below is a satellite picture of one, courtesy of Arctic Climatology and Meteorology. I think you will agree it looks most impressive. One can almost ‘see’ the snow showers falling under those large white clumps of cloud! A NOAA-9 polar orbiter satellite image (visible band) of a polar low over the Barents Sea on 27 February 1987. The southern tip of Spitsbergen is visible at the top of the image. The polar low is centered just north of the Norwegian coast. Image contributed by S. Businger, Department of Meteorology, University of Hawaii. A Polar Low is what is termed a mesoscale feature. By that it is a relatively small feature within the size of normal depression types. Covering hundreds of square kilometers not thousands as the main Polar Front Depressions do. It only occurs in winter time but otherwise has many characteristics of a ‘comma cloud’. This in a simple explanation is a mass of deep convective cloud which tends to organize itself into a comma shape. Its formation is as a ‘heat’ low! This may sound odd in mid winter but its the heat of the sea and the very cold air over the top that causes the Polar Low to form. It can even, on a satellite picture, give the appearance of a hurricane, not that it is of course. Although strong winds can occur in its vicinity producing blizzard conditions. Please forgive me if I keep things as simple as I possibly can. I do this in any teaching session. That way more people come to understand the basics and others can ask questions at the end of the item or read the various other links. Polar Lows tend to form, in the area of our interest, in a north or north west airflow. It is possible for it to occur in the Norwegian Sea in a north east flow but is less common. One area they seem to favour is just south of Iceland. Quite why is open to discussion. It may be that some kind of ‘lee’ effect is the first trigger. By that I mean air flowing over the mountains of Iceland then coming over markedly warmer sea to the south. The two combine and a polar low is formed. They will normally form in areas of relatively lightish winds, sometimes on the edge of a strong surface flow. Again some kind of lee or eddy effect may help this formation. Initially the cloud is all convective, Cumulus and Cumulonimbus, but as the feature develops middle and sometimes high level cloud will be created. As the feature moves further south its source of energy is the sea. Coming south it encounters increasingly warmer sea thus aiding more instability and more convective cloud along with all the moisture it needs from the sea. It usually, not always, tends to track in the overall direction of the isobars. Another thing to look at is the skew-T diagram and the winds at 5,000ft and 10,000 ft also give it some of its direction. Once near the British Isles it will have a tendency to try and stay over water whilst it can. So the entrance to the North Channel between Ireland and Scotland is quite a favourite. Trying to predict its track is not easy. But following the suggestions above will give good guidance. Another thing to look for are pressure falls, as with all lows, they tend to move in the direction of the largest fall of pressure. Indeed if it is possible, and this is not always so due to sparse data over the north east Atlantic. However, if it is possible then drawing the area of higher pressure rises behind it with the higher falls ahead of it, then joining this imaginary line can give quite a good idea of its steering over 3-6 hours. Will it snow or not? The age old question. So what, if we have a Polar Low approaching, are the things to look for. 1) time of day 2) link to 1) with temperature 3) the dewpoint 4) freezing level(0 deg isotherm) There are plenty more but these are readily available to anyone on Net Weather and will give a fair guide to what may happen. Look at its track, using the Met Office Fax charts as they come out. Make a rough diagram of the area between Iceland and the UK and plot its movement each 6 hours. (The map below may help). This will give a good indication of its likely track for the next 3-6 hours or so. By which time the next Fax chart should be available. Watch out for your Net Wx forecast team issuing forecasts or Alerts! Below is a map, with acknowledgement which might help. It has the area needed, latitude and longitude and a scale. I hope it helps. A more technical definition by Rasmussen and Turner (2003) states: 'A polar low is a small, but fairly intense maritime cyclone that forms poleward of the main baroclinic zone (the polar front or other major baroclinic zone). The horizontal scale of the polar low is approximately between 200 and 1000 kilometres and surface winds near or above gale force.' This shows how rough at sea it can be near a Polar Low. John Holmes
    1 point
  7. Here are the current Papers & Articles under the research topic Madden-Julian Oscillation (MJO). Click on the title of a paper you are interested in to go straight to the full paper. Papers and articles covering the basics (ideal for learning) are shown in Green. What is the MJO, and why do we care? No abstract, so a snippet: Imagine ENSO as a person riding a stationary exercise bike in the middle of a stage all day long. His unchanging location is associated with the persistent changes in tropical rainfall and winds that we have previously described as being linked to ENSO. Now imagine another bike rider entering the stage on the left and pedaling slowly across the stage, passing the stationary bike (ENSO), and exiting the stage at the right. This bike rider we will call the MJO and he/she may cross the stage from left to right several times during the show. So, unlike ENSO, which is stationary, the MJO is an eastward moving disturbance of clouds, rainfall, winds, and pressure that traverses the planet in the tropics and returns to its initial starting point in 30 to 60 days, on average. Learning All About the MJO No abstract, but this fascinating article mainly focusses on how the MJO impact spreads globally and includes some great illustrations. Fifty Years of Research on the Madden-Julian Oscillation: Recent Progress, Challenges, and Perspectives Published July 2020. Abstract: Since its discovery in the early 1970s, the crucial role of the Madden-Julian Oscillation (MJO) in the global hydrological cycle and its tremendous influence on high-impact climate and weather extremes have been well recognized. The MJO also serves as a primary source of predictability for global Earth system variability on subseasonal time scales. The MJO remains poorly represented in our state-of-the-art climate and weather forecasting models, however. Moreover, despite the advances made in recent decades, theories for the MJO still disagree at a fundamental level. The problems of understanding and modeling the MJO have attracted significant interest from the research community. As a part of the AGU's Centennial collection, this article provides a review of recent progress, particularly over the last decade, in observational, modeling, and theoretical study of the MJO. A brief outlook for near-future MJO research directions is also provided. Improving the prediction of the Madden-Julian Oscillation of the ECMWF model by post-processing Published: March 2022 Abstract: The Madden-Julian Oscillation (MJO) is a major source of predictability on the sub-seasonal (10- to 90-days) time scale. An improved forecast of the MJO, may have important socioeconomic impacts due to the influence of MJO on both, tropical and extratropical weather extremes. Although in the last decades state-of-the-art climate models have proved their capability for forecasting the MJO exceeding the 5 weeks prediction skill, there is still room for improving the prediction. In this study we use Multiple Linear Regression (MLR) and a Machine Learning (ML) algorithm as post-processing methods to improve the forecast of the model that currently holds the best MJO forecasting performance, the European Centre for Medium-Range Weather Forecast (ECMWF) model. We find that both MLR and ML improve the MJO prediction and that ML outperforms MLR. The largest improvement is in the prediction of the MJO geographical location and intensity. Deep learning for bias correction of MJO prediction Published: May 2021 Abstract: Producing accurate weather prediction beyond two weeks is an urgent challenge due to its ever-increasing socioeconomic value. The Madden-Julian Oscillation (MJO), a planetary-scale tropical convective system, serves as a primary source of global subseasonal (i.e., targeting three to four weeks) predictability. During the past decades, operational forecasting systems have improved substantially, while the MJO prediction skill has not yet reached its potential predictability, partly due to the systematic errors caused by imperfect numerical models. Here, to improve the MJO prediction skill, we blend the state-of-the-art dynamical forecasts and observations with a Deep Learning bias correction method. With Deep Learning bias correction, multi-model forecast errors in MJO amplitude and phase averaged over four weeks are significantly reduced by about 90% and 77%, respectively. Most models show the greatest improvement for MJO events starting from the Indian Ocean and crossing the Maritime Continent. ENSO Modulation of MJO Teleconnections to the North Atlantic and Europe Published Nov 2019 Abstract: The teleconnection from the Madden-Julian Oscillation (MJO) provides a source of subseasonal variability and predictability to the North Atlantic-European (NAE) region. The El Niño-Southern Oscillation (ENSO) modulates the seasonal mean state, through which the MJO and its teleconnection pattern propagates; however, its impact on this teleconnection to the NAE region has not been investigated. Here we find a robust dependence of the teleconnections from the MJO to NAE weather regimes on the phase of ENSO. We show that the MJO to NAO+ regime tropospheric teleconnection is strongly enhanced during El Niño years, via enhanced Rossby wave activity, and suppressed during La Niña. Conversely, the MJO to NAO− regime stratospheric teleconnection is enhanced during La Niña years and suppressed during El Niño. This dependence on the background state has strong implications for subseasonal predictability, including interannual variations in subseasonal predictive skill. Verification of medium-range MJO forecasts with TIGGE Published June 2011 Abstract: The Madden-Julian oscillation (MJO) is the dominant mode of intraseasonal variability in the tropics. Accurate simulations of the MJO are important for studies of weather and climate in the tropics and extratropics. This study assesses the forecast performance of operational medium-range ensemble forecasts, available at THe Observing system Research and Predictability EXperiment (THORPEX) Interactive Grand Global Ensemble (TIGGE) data portal, regarding the MJO for the past 3 years. The results indicate that ECMWF (European Centre for Medium-range Weather Forecasts) and UKMO (United Kingdom Meteorological Office) generally yield the best performances in predicting the MJO; however, they do not always show similar skills. ECMWF performs well in simulating the maintenance and onset of the MJO in phases 1-4, whereas UKMO and NCEP (National Centers for Environmental Prediction) perform well in simulating the maintenance and onset of the MJO in phases 5-8. Thus, the best-performing numerical weather prediction (NWP) centre varies with the phase of the MJO. With advance knowledge of the forecast characteristics of each NWP centre, we can ensure more reliable forecasts of the MJO in operational uses, based on the MJO phase. This represents an advantage of the multi-centre grand ensemble approach. The Role of MJO Propagation, Lifetime, and Intensity on Modulating the Temporal Evolution of the MJO Extratropical Response Published May 2019 Abstract: The Madden-Julian Oscillation (MJO) is the dominant mode of tropical intraseasonal variability. Many studies have found that the MJO has significant impacts on extratropical weather. Since the MJO can act as a tropical heat source that excites Rossby waves, midlatitude weather is modulated by the MJO due to the Rossby waves that propagate into the midlatitude and modulate the midlatitude circulation. Heat sources of individual MJO events are different since each event has different eastward propagation speed, lifetime, intensity, and structure. The background flow is also different for each event. These result in different Rossby waves and different extratropical response for each MJO event. In this study, the role of MJO propagation speed, lifetime, and intensity on modulating the structure and temporal evolution of the MJO extratropical response is systematically explored by using an idealized general circulation model. By adding the MJO-associated heating into the general circulation model as an external forcing, the extratropical response in the Reanalysis is captured reasonably by the model. However, large ensemble model simulations show that the response in the Reanalysis is not robust. Experiments with MJO events of different propagation speed, lifetime, and intensity show that to excite a strong extratropical response, the MJO has to propagate through specific phases (Phases 1–3 and 5–7). The intensity, timing, and duration of the extratropical response strongly depend on when the MJO is initiated and when the MJO decays. The extratropical impacts of slow- and fast-propagating MJO also have significant differences, especially on intensity and duration. Changes in Madden‐Julian Oscillation Precipitation and Wind Variance Under Global Warming Published Jul 2018. Abstract: The Madden‐Julian oscillation (MJO) is the leading mode of tropical intraseasonal variability, having profound impacts on many weather and climate phenomena across the tropics and extratropics. Previous studies using a limited number of models have suggested complex changes in MJO activity in a warmer climate. While most studies have argued that MJO precipitation amplitude will increase in a future warmer climate, others note that this is not necessarily the case for MJO wind variability. This distinction is important since MJO wind fluctuations are responsible for producing remote impacts on extreme weather through teleconnections. In this study, we examine projected changes of MJO precipitation and wind variance at the end of the 21st century in Representative Concentration Pathway 8.5 using the multimodel Coupled Model Intercomparison Project phase 5 data set. Under global warming, most models show an increase in MJO band precipitation variance, while wind variability decreases. The discrepancy between MJO precipitation and wind variance changes under global warming is shown to be due to increases in tropical static stability in a warmer climate. The multimodel mean shows a 20% increase in both the 500‐hPa vertical tropical dry static energy gradient and the ratio of intraseasonal precipitation to 500 hPa omega fluctuations, consistent with scaling by weak temperature gradient theory. These results imply that tropical static stability increases may weaken the MJO's ability to influence extreme events in future warmer climate by weakening wind teleconnections, even though MJO precipitation amplitude may increase. A Cautionary Note on the Long‐term Trend in Activity of the Madden‐Julian Oscillation During the Past Decades Published Dec 2019. Abstract: Recent studies suggest that frequency of active phases of the Madden‐Julian Oscillation (MJO) over the Maritime Continent and western Pacific, that is, the MJO Phases 4–6 defined by the real‐time multivariate MJO (RMM) index, has increased in recent winters. A robust positive trend in MJO Phase 4–6 days during 1979–2015 winters is confirmed in this study. Our analyses, however, suggest that this trend could be exaggerated due to the blended low‐frequency variability signals in the RMM. When the winter RMM is reconstructed using anomalous fields after removing their winter mean instead of the previous 120‐day mean as for the original RMM, the robust trend in MJO Phase 4–6 days can no longer be detected. Therefore, cautions need to be exercised when applying the RMM for studies on the low‐frequency variability and climate trend in MJO activity and using the derived MJO variability to interpret associated changes in climate systems. A Madden‐Julian Oscillation event remotely accelerates ocean upwelling to abruptly terminate the 1997/1998 super El Niño Abstract: The termination of the super intense 1997/1998 El Niño was extraordinarily abrupt. The May 1998 Madden-Julian Oscillation (MJO), a massive complex of stormy tropical clouds, is among possible contributors to the abrupt termination. Despite having been sensationally proposed 18 years ago, the role of the MJO remained controversial and speculative because of the difficulty of sufficiently simulating the El Niño and MJO simultaneously. An ensemble simulation series using a newly developed, fully ocean-coupled version of a global cloud system resolving numerical model replicated the specific atmosphere and ocean conditions of May 1998 in unprecedented detail, extending the prediction skill of the MJO to 46 days. Simulation ensemble members with stronger MJO activities over the Maritime Continent experienced quicker sea surface temperature drop in the eastern Pacific, confirming that the easterly winds associatedwith the remote MJO accelerated ocean upwelling to abruptly terminate the El Niño. An All-Season Real-Time Multivariate MJO Index: Development of an Index for Monitoring and Prediction Abstract: A seasonally independent index for monitoring the Madden–Julian oscillation (MJO) is described. It is based on a pair of empirical orthogonal functions (EOFs) of the combined fields of near-equatorially averaged 850-hPa zonal wind, 200-hPa zonal wind, and satellite-observed outgoing longwave radiation (OLR) data. Projection of the daily observed data onto the multiple-variable EOFs, with the annual cycle and components of interannual variability removed, yields principal component (PC) time series that vary mostly on the intraseasonal time scale of the MJO only. This projection thus serves as an effective filter for the MJO without the need for conventional time filtering, making the PC time series an effective index for real-time use. The pair of PC time series that form the index are called the Real-time Multivariate MJO series 1 (RMM1) and 2 (RMM2). The properties of the RMM series and the spatial patterns of atmospheric variability they capture are explored. Despite the fact that RMM1 and RMM2 describe evolution of the MJO along the equator that is independent of season, the coherent off-equatorial behavior exhibits strong seasonality. In particular, the northward, propagating behavior in the Indian monsoon and the southward extreme of convection into the Australian monsoon are captured by monitoring the seasonally independent eastward propagation in the equatorial belt. The previously described interannual modulation of the global variance of the MJO is also well captured. Applications of the RMM series are investigated. One application is through their relationship with the onset dates of the monsoons in Australia and India; while the onsets can occur at any time during the convectively enhanced half of the MJO cycle, they rarely occur during the suppressed half. Another application is the modulation of the probability of extreme weekly rainfall; in the “Top End” region around Darwin, Australia, the swings in probability represent more than a tripling in the likelihood of an upper-quintile weekly rainfall event from the dry to wet MJO phase. Circulation Response to Fast and Slow MJO Episodes 2017 paper. Abstract: Fast and slow Madden–Julian oscillation (MJO) episodes have been identified from 850- and 200-hPa zonal wind and outgoing longwave radiation (OLR) for 32 winters (16 October–17 March) 1980/81–2011/12. For 26 fast cases the OLR took no more than 10 days to propagate from phase 3 (convection over the Indian Ocean) to phase 6 (convection over the western Pacific). For 8 slow cases the propagation took at least 20 days. Fast episode composite anomalies of 500-hPa height (Z500) show a developing Rossby wave in the mid-Pacific with downstream propagation through MJO phases 2–4. Changes in the frequency of occurrence of the NAO1 weather regime are modest. This Rossby wave is forced by anomalous cooling over the Maritime Continent during phases 2 and 3 (seen in phase-independent wave activity flux). The upper-level anticyclonic response to phase-3 heating is a secondary source of wave activity. The Z500 slow episode composite response to MJO phases 1 and 2 is an enhanced Aleutian low followed by a North American continental high. Following phase 4 the development of an NAO1 like pattern is seen over the Atlantic, transitioning to a strong NAO2 pattern by phase 8. A dramatic increase in frequency of the NAO1 weather regime follows phases 4and 5, while a strong increase in NAO2 regime follows phases 6 and 7. The responses to MJO-related heating and cooling over the Indian and western Pacific Oceans in phases 1–4 provide a source for wave activity propagating to North America, augmented by storm-track anomalies. Combined effect of the QBO and ENSO on the MJO Abstract: This study investigates the combined effect of the El Niño–Southern Oscillation (ENSO) and stratospheric quasi-biennial oscillation (QBO) on the Madden Julian Oscillation (MJO). Here, the authors investigate the combined effect of the QBO and ENSO on the MJO. The results show that the western Pacific MJO originating from the Indian Ocean during La Niña/QBO easterly years is stronger than that during El Niño years. This relation, however, disappears during La Niña/QBO westerly years. The reason is that ENSO and the QBO have different effects on each MJO event. For an El Niño year, there is only about one MJO event, and the QBO effect is small. During a La Niña/QBO easterly year, there are 1.7 MJO events, while during a La Niña/QBO westerly year, there are only 0.6 MJO events. El Niño can reinforce the MJO over the western Pacific because of the positive moisture advection of the El Niño mean state by MJO easterly wind anomalies. The QBO mainly affects the MJO over the Maritime Continent region by changing the high-cloud-controlled diurnal cycle; and the Maritime Continent barrier effect is enhanced during the QBO westerly phase because of the strong diurnal cycle. During El Niño years, even the MJO over the Maritime Continent is suppressed by the QBO westerly phase; the MJO can be reinforced over the western Pacific. During La Niña/QBO westerly years, the MJO over the Maritime Continent is suppressed because of the strong Maritime Continent diurnal cycle, and it is further suppressed over the western Pacific because of the lack of a reinforcement process. Impact of the MJO on the boreal winter extratropical circulation Published August 2014. Abstract: The effect of the Madden‐Julian Oscillation (MJO) on the Northern Hemisphere wintertime stratospheric polar vortex is evaluated using a meteorological reanalysis data set and a modern atmospheric general circulation model. The MJO influences the tropospheric North Pacific; and in particular, it modulates the heat flux that is in phase with the climatological planetary waves in both the troposphere and stratosphere. The phase of the MJO in which convection is propagating into the tropical central Pacific leads to a weakened vortex, while suppressed MJO convection in this region is associated with a stronger vortex. Subsequently, the MJO modulates the phase of the tropospheric North Atlantic Oscillation (also known as the Arctic Oscillation or the Northern Annular Mode). While the responses in the model and in the reanalysis data differ in some respects, they both indicate that the MJO can remotely impact the extratropical tropospheric circulation via the stratosphere. Influence of the QBO on MJO prediction skill in the subseasonal-to-seasonal prediction models Abstract: Recent studies have shown that the Madden-Julian Oscillation (MJO) is significantly modulated by the stratospheric Quasi Biennial Oscillation (QBO). In general, boreal winter MJO becomes more active during the easterly phase of the QBO (EQBO) than during the westerly phase (WQBO). Based on this finding, here we examine the possible impacts of the QBO on MJO prediction skill in the operational models that participated in the WCRP/WWRP subseasonal-to-seasonal (S2S) prediction project. All models show a higher MJO prediction skill during EQBO winters than during WQBO winters. For the bivariate anomaly correlation coefficient of 0.5, the MJO prediction skill during EQBO winters is enhanced by up to 10 days. This enhancement is insensitive to the initial MJO amplitude, indicating that the improved MJO prediction skill is not simply the result of a stronger MJO. Instead, a longer persistence of the MJO during EQBO winters likely induces a higher prediction skill by having a higher prediction limit. Influence of the Stratospheric Quasi-Biennial Oscillation on the Madden–Julian Oscillation during Austral Summer Abstract: Influence of the stratospheric quasi-biennial oscillation (QBO) on the Madden–Julian oscillation (MJO) and its statistical significance are examined for austral summer (DJF) in neutral ENSO events during 1979–2013. The amplitude of the OLR-based MJO index (OMI) is typically larger in the easterly phase of the QBO at 50 hPa (E-QBO phase) than in the westerly (W-QBO) phase. Daily composite analyses are performed by focusing on phase 4 of the OMI, when the active convective system is located over the eastern Indian Ocean through the Maritime Continent. The composite OLR anomaly shows a larger negative value and slower eastward propagation with a prolonged period of active convection in the E-QBO phase than in the W-QBO phase. Statistically significant differences of the MJO activities between the QBO phases also exist with dynamical consistency in the divergence of horizontal wind, the vertical wind, the moisture, the precipitation, and the 100-hPa temperature. A conditional sampling analysis is also performed by focusing on the most active convective region for each day, irrespective of the MJO amplitude and phase. Composite vertical profiles of the conditionally sampled data over the most active convective region reveal lower temperature and static stability around the tropopause in the E-QBO phase than in the W-QBO phase, which indicates more favorable conditions for developing deep convection. This feature is more prominent and extends into lower levels in the upper troposphere over the most active convective region than other tropical regions. Composite longitude–height sections show similar features of the large-scale convective system associated with the MJO, including a vertically propagating Kelvin response. Intraseasonal interaction between the Madden–Julian Oscillation and the North Atlantic Oscillation 2008 paper. Abstract: Bridging the traditional gap between the spatio-temporal scales of weather and climate is a significant challenge facing the atmospheric community. In particular, progress in both medium-range and seasonal-to-interannual climate prediction relies on our understanding of recurrent weather patterns and the identification of specific causes responsible for their favoured occurrence, persistence or transition. Within this framework, I here present evidence that the main climate intra-seasonal oscillation in the tropics—the Madden–Julian Oscillation1,2 (MJO)—controls part of the distribution and sequences of the four daily weather regimes defined over the North Atlantic–European region in winter3. North Atlantic Oscillation4 (NAO) regimes are the most affected, allowing for medium-range predictability of their phase far exceeding the limit of around one week that is usually quoted. The tropical–extratropical lagged relationship is asymmetrical. Positive NAO events mostly respond to a mid-latitude low-frequency wave train initiated by the MJO in the western–central tropical Pacific and propagating eastwards. Precursors for negative NAO events are found in the eastern tropical Pacific–western Atlantic, leading to changes along the North Atlantic storm track. Wave-breaking diagnostics tend to support the MJO preconditioning and the role of transient eddies in setting the phase of the NAO. I present a simple statistical model to quantitatively assess the potential predictability of the daily NAO index or the sign of the NAO regimes when they occur. Forecasts are successful in ∼70 per cent of the cases based on the knowledge of the previous ∼12-day MJO phase used as a predictor. This promising skill could be of importance considering the tight link4 between weather regimes and both mean conditions and the chances of extreme events occurring over Europe. These findings are useful for further stressing the need to better simulate and forecast the tropical coupled ocean–atmosphere dynamics, which is a source of medium-to-long range predictability and is the Achilles’ heel of the current seamless prediction suites. Circulation Response to Fast and Slow MJO Episodes 2008 paper. Abstract: In this study, we detected the spatial and temporal characteristics of Madden-Julian Oscillation (MJO) using zonal winds at the surface and outgoing long-wave radiation (OLR) from the NCEP-NCAR (U.S. National Center of Environmental Prediction-National Center for Atmospheric Research) reanalysis product from 1981–2003. The results show that MJO activity, represented by these two variables, has large variances around 10° off the equator and over the near-equatorial western Pacific. One central issue addressed in this study is MJO-ENSO (El Niño and Southern Oscillation) relationship. It has been found that there exists a statistically significant relationship between MJO in spring-summer and ENSO in autumn-winter. The relationship of MJO-ENSO is nonlinear in nature and has a decadal variation. A much stronger statistical relationship of MJO-ENSO was found in the 1990s as compared to that in the 1980s. These findings were further verified using ECMWF (European Center for Medium-Range Weather Forecasts) reanalysis product. The potential mechanisms responsible for MJO-ENSO relationship are also discussed. MJO‐Related Tropical Convection Anomalies Lead to More Accurate Stratospheric Vortex Variability in Sub-seasonal Forecast Models Abstract: The effect of the Madden‐Julian Oscillation (MJO) on the Northern Hemisphere wintertime stratospheric polar vortex in the period preceding stratospheric sudden warmings is evaluated in operational subseasonal forecasting models. Reforecasts which simulate stronger MJO‐related convection in the Tropical West Pacific also simulate enhanced heat flux in the lowermost stratosphere and a more realistic vortex evolution. The time scale on which vortex predictability is enhanced lies between 2 and 4 weeks for nearly all cases. Those stratospheric sudden warmings that were preceded by a strong MJO event are more predictable at ∼20 day leads than stratospheric sudden warmings not preceded by a MJO event. Hence, knowledge of the MJO can contribute to enhanced predictability, at least in a probabilistic sense, of the Northern Hemisphere polar stratosphere. Modulation of equatorial Pacific westerly/easterly wind events by the Madden–Julian oscillation and convectively‑coupled Rossby waves Abstract: Synoptic wind events in the equatorial Pacific strongly influence the El Niño/Southern Oscillation (ENSO) evolution. This paper characterizes the spatio-temporal distribution of Easterly (EWEs) and Westerly Wind Events (WWEs) and quantifies their relationship with intraseasonal and interannual large-scale climate variability. We unambiguously demonstrate that the Madden–Julian Oscillation (MJO) and Convectively-coupled Rossby Waves (CRW) modulate both WWEs and EWEs occurrence probability. 86 % of WWEs occur within convective MJO and/or CRW phases and 83 % of EWEs occur within the suppressed phase of MJO and/or CRW. 41 % of WWEs and 26 % of EWEs are in particular associated with the combined occurrence of a CRW/MJO, far more than what would be expected from a random distribution (3 %). Wind events embedded within MJO phases also have a stronger impact on the ocean, due to a tendency to have a larger amplitude, zonal extent and longer duration. These findings are robust irrespective of the wind events and MJO/CRW detection methods. While WWEs and EWEs behave rather symmetrically with respect to MJO/CRW activity, the impact of ENSO on wind events is asymmetrical. The WWEs occurrence probability indeed increases when the warm pool is displaced eastward during El Niño events, an increase that can partly be related to interannual modulation of the MJO/CRW activity in the western Pacific. On the other hand, the EWEs modulation by ENSO is less robust, and strongly depends on the wind event detection method. The consequences of these results for ENSO predictability are discussed. Modulation of the Global Atmospheric Circulation by Combined Activity in the Madden–Julian Oscillation and the El Niño–Southern Oscillation during Boreal Winter 2010 paper. Abstract: Composite global patterns associated with the El Niño–Southern Oscillation (ENSO) and the Madden–Julian oscillation (MJO) are frequently applied to help make predictions of weather around the globe at lead times beyond a few days. However, ENSO modulates the background states through which the MJO and its global response patterns propagate. This paper explores the possibility that nonlinear variations confound the combined use of composites based on the MJO and ENSO separately. Results indicate that when both modes are active at the same time, the associated patterns in the global flow are poorly represented by simple linear combinations of composites based on the MJO and ENSO individually. Composites calculated by averaging data over periods when both modes are present at the same time more effectively describe the associated weather patterns. Results reveal that the high-latitude response to the MJO varies with ENSO over all longitudes, but especially across the North Pacific Rim, North America, and the North Atlantic. Further analysis demonstrates that the MJO influence on indexes of the North Atlantic Oscillation is greatest during La Niña conditions or during periods of rapid adjustment in the phase of ENSO. More Frequent Sudden Stratospheric Warming Events due to Enhanced MJO Forcing Expected in a Warmer Climate Abstract: Sudden stratospheric warming (SSW) events influence the Arctic Oscillation and midlatitude extreme weather. Observations show SSW events to be correlated with certain phases of the Madden–Julian oscillation (MJO), but the effect of the MJO on SSW frequency is unknown, and the teleconnection mechanism, its planetary wave propagation path, and time scale are still not completely understood. The Arctic stratosphere response to increased MJO forcing expected in a warmer climate using two models is studied: the comprehensive Whole Atmosphere Community Climate Model and an idealized dry dynamical core with and without MJO-like forcing. It is shown that the frequency of SSW events increases significantly in response to stronger MJO forcing, also affecting the averaged polar cap temperature. Two teleconnection mechanisms are identified: a direct propagation of MJO-forced transient waves to the Arctic stratosphere and a nonlinear enhancement of stationary waves by the MJO-forced transient waves. The MJO-forced waves propagate poleward in the lower stratosphere and upper troposphere and then upward. The cleaner results of the idealized model allow identifying the propagating signal and suggest a horizontal propagation time scale of10–20 days, followed by additional time for upward propagation within the Arctic stratosphere, although there are significant uncertainties involved. Given that the MJO is predicted to be stronger in a warmer climate, these results suggest that SSW events may become more frequent, with possible implications on tropospheric high-latitude weather. However, the effect of an actual warming scenario on SSW frequency involves additional effects besides a strengthening of the MJO, requiring further investigation. Northern Hemisphere mid‐winter vortex‐displacement and vortex‐split stratospheric sudden warmings: Influence of the Madden‐Julian Oscillation and Quasi‐Biennial Oscillation Abstract: We investigate the connection between the equatorial Madden-Julian Oscillation (MJO) and different types of the Northern Hemisphere mid-winter major stratospheric sudden warmings (SSWs), i.e., vortex-displacement and vortex-split SSWs. The MJO-SSW relationship for vortex-split SSWs is stronger than that for vortex-displacement SSWs, as a result of the stronger and more coherent eastward propagating MJOs before vortex-split SSWs than those before vortex-displacement SSWs. Composite analysis indicates that both the intensity and propagation features of MJO may influence the MJO-related circulation pattern at high latitudes and the type of SSWs. A pronounced Quasi-Biennial Oscillation (QBO) dependence is found for vortex-displacement and vortex-split SSWs, with vortex-displacement (-split) SSWs occurring preferentially in easterly (westerly) QBO phases. The lagged composites suggest that the MJO-related anomalies in the Arctic are very likely initiated when the MJO-related convection is active over the equatorial Indian Ocean (around the MJO phase 3). Further analysis suggests that the QBO may modulate the MJO-related wave disturbances via its influence on the upper tropospheric subtropical jet. As a result, the MJO-related circulation pattern in the Arctic tends to be wave number-one/wave number-two ~25–30 days following phase 3 (i.e., approximately phases 7–8, when the MJO-related convection is active over the western Pacific) during easterly/westerly QBO phases, which resembles the circulation pattern associated with vortex-displacement/vortex-split SSWs. Observed Changes in the Lifetime and Amplitude of the MJO Associated with Interannual ENSO Sea Surface Temperature Anomalies Abstract: The Madden–Julian oscillation (MJO) is analyzed using the reanalysis zonal wind– and satellite outgoing longwave radiation–based indices of Wheeler and Hendon for the 1974–2005period. The average lifetime of the MJO events varies with season (36 days for events whose central date occurs in December, and 48days for events in September). The lifetime of the MJO in the equinoctial seasons (March–May andOctober–December) is also dependent on the state of El Niño–Southern Oscillation (ENSO). During October–December it is only 32 days under El Niño conditions, increasing to 48 days under La Niña conditions, with similar values in northern spring. This difference is due to faster eastward propagation of the MJO convective anomalies through the Maritime Continent and western Pacific during El Niño,consistent with theoretical arguments concerning equatorial wave speeds.The analysis is extended back to 1950 by using an alternative definition of the MJO based on just the zonal wind component of the Wheeler and Hendon indices. A rupture in the amplitude of the MJO is foundin 1975, which is at the same time as the well-known rupture in the ENSO time series that has been associated with the Pacific decadal oscillation. The mean amplitude of the MJO is 16% larger in the postrupture (1976–2005) compared to the prerupture (1950–75) period. Before the 1975 rupture, the amplitude of the MJO is maximum (minimum) under El Niño (La Niña) conditions during northern winter,and minimum (maximum) under El Niño (La Niña) conditions during northern summer. After the rupture,this relationship disappears. When the MJO–ENSO relationship is analyzed using all-year-round data, or ashorter dataset (as in some previous studies), no relationship is found. Observed connection between stratospheric sudden warmings and the Madden-Julian Oscillation Abstract: The effect of the Madden-Julian Oscillation (MJO) on the Northern Hemisphere wintertime stratospheric polar vortex and major, mid-winter stratospheric sudden warmings (SSWs) is evaluated using a meteorological reanalysis dataset. The MJO influences the region in the tropospheric North Pacific sector that is most strongly associated with a SSW.Consistent with this, SSWs in the reanalysis record have tended to follow certain MJO phases. The magnitude of the influence of the MJO on the vortex is comparable to that associated with the Quasi-Biennial Oscillation and El Niño. The MJO could be used to improve intra-seasonal projections of the Northern Hemisphere high latitude circulation, and in particular of the tropospheric Northern Annular Mode, at lags exceeding one month. Ocean Rossby waves as a triggering mechanism for primary Madden-Julian events Abstract: The Madden–Julian Oscillation (MJO) is sporadic, with episodes of cyclical activity interspersed with inactive periods. However, it remains unclear what may trigger a Madden–Julian (MJ) event which is not immediately preceded by any MJO activity:a ‘primary’ MJ event. A combination of case-studies and composite analysis is usedto examine the extent to which the triggering of primary MJ events might occur in response to ocean dynamics. The case-studies show that such events can be triggered by the arrival of a downwelling oceanic equatorial Rossby wave, which is shown to be associated with a deepening of the mixed layer and positive sea-surface temperature(SST) anomalies of the order of 0.5–1◦C. These SST anomalies are not attributable to forcing by surface fluxes which are weak for the case-studies analysed. Furthermore,composite analysis suggests that such forcing is consistently important for triggering primary events. The relationship is much weaker for successive events, due to themany other triggering mechanisms which operate during periods of cyclical MJO activity. This oceanic feedback mechanism is a viable explanation for the sporadic and broadband nature of the MJO. Additionally, it provides hope for forecasting MJ events during periods of inactivity, when MJO forecasts generally exhibit low skill. On the emerging relationship between the stratospheric Quasi-Biennial oscillation and the Madden-Julian oscillation Abstract: A strong relationship between the quasi-biennial oscillation (QBO) of equatorial stratospheric winds and the amplitude of the Madden-Julian oscillation (MJO) during the boreal winter has recently been uncovered using observational data from the mid-1970s to the present. When the QBO is in its easterly phase in the lower stratosphere, it favors stronger MJO activity during boreal winter, while the MJO tends to be weaker during the westerly phase of the QBO. Here we show using reconstructed indices of the MJO and QBO back to 1905 that the relationship between enhanced boreal winter MJO activity and the easterly phase of the QBO has only emerged since the early 1980s. The emergence of this relationship coincides with the recent cooling trend in the equatorial lower stratosphere and the warming trend in the equatorial upper troposphere, which appears to have sensitized MJO convective activity to QBO-induced changes in static stability near the tropopause. Climate change is thus suggested to have played a role in promoting coupling between the MJO and the QBO. Predictability and Prediction of the North Atlantic Oscillation (large section on MJO impact) 2010 article. No abstract, but the introduction: The North Atlantic Oscillation (NAO) is one of the most important modes of variability in the Northern Hemisphere extratropical atmosphere. The NAO has a wide range of time scales from days to decades. It has larger amplitude in boreal winter than in summer. On interannual time scale, the NAO accounts for about 30% of the hemispheric surface air temperature variance over the past 60 winters. The NAO has long been recognized as a major circulation pattern influencing the weather and climate from eastern North America to Europe. It is generally accepted that the primary mechanism for the NAO is the internal dynamics of the extratropical circulation. This implies a lack of predictability for the NAO variability beyond the time scale of a baroclinic wave. Processes external to the extratropical atmosphere have also been found to contribute to the NAO variability on intraseasonal and seasonal scales. For example, tropical diabatic heating anomalies associated with the Madden-Julian Oscillation (MJO) can influence the extratropical circulation and the NAO. Sea surface temperature (SST) and snow cover anomalies may have an impact on the NAO on seasonal time scale. In this article, focus of discussion is made on the forecast skill of the NAO on intraseasonal and seasonal time scales. QBO Influence on MJO Amplitude over the Maritime Continent:Physical Mechanisms and Seasonality Abstract: The quasi-biennial oscillation (QBO) is stratified by stratospheric zonal wind direction and height intofour phase pairs [easterly midstratospheric winds (QBOEM), easterly lower-stratospheric winds, westerly midstratospheric winds (QBOWM), and westerly lower-stratospheric winds] using an empirical orthogonal function analysis of daily stratospheric (100–10 hPa) zonal wind data during 1980–2017. Madden–Julian oscillation (MJO) events in which the MJO convective envelope moved eastward across the Maritime Continent (MC) during 1980–2017 are identified using the Real-time Multivariate MJO (RMM) index and theoutgoing longwave radiation (OLR) MJO index (OMI). Comparison of RMM amplitudes by the QBO phase pair over the MC (RMM phases 4 and 5) reveals that boreal winter MJO events have the strongest amplitudes during QBOEM and the weakest amplitudes during QBOWM, which is consistent with QBO-driven differences in upper-tropospheric lower-stratospheric (UTLS) static stability. Additionally, boreal winter RMM events over the MC strengthen during QBOEM and weaken during QBOWM. In the OMI, those amplitude changes generally shift eastward to the eastern MC and western Pacific Ocean, which may result from differences in RMM and OMI index methodologies. During boreal summer, as the northeastward-propagating boreal summer intraseasonal oscillation (BSISO) becomes the dominant mode of intraseasonal variability, these relationships are reversed. Zonal differences in UTLS stability anomalies are consistent with amplitude changes of eastward-propagating MJO events across the MC during boreal winter, and meridional stability differences are consistent with amplitude changes of northeastward-propagating BSISO events during boreal summer. Results remain consistent when stratifying by neutral ENSO phase. Stratospheric Control of the Madden–Julian Oscillation Abstract: Interannual variation of seasonal-mean tropical convection over the Indo-Pacific region is primarily controlled by El Niño–Southern Oscillation (ENSO). For example, during El Niño winters, seasonal-mean convection around the Maritime Continent becomes weaker than normal, while that over the central to eastern Pacific is strengthened. Similarly, subseasonal convective activity, which is associated with the Madden–Julian oscillation (MJO), is influenced by ENSO. The MJO activity tends to extend farther eastward to the date line during El Niño winters and contract toward the western Pacific during La Niña winters. However, the overall level of MJO activity across the Maritime Continent does not change much in response to the ENSO. It is shown that the boreal winter MJO amplitude is closely linked with the stratospheric quasi-biennial oscillation (QBO) rather than with ENSO. The MJO activity around the Maritime Continent becomes stronger and more organized during the easterly QBO winters. The QBO-related MJO change explains up to 40% of interannual variation of the boreal winter MJO amplitude. This result suggests that variability of the MJO and the related tropical–extratropical teleconnections can be better understood and predicted by taking not only the tropospheric circulation but also the stratospheric mean state into account. The seasonality of the QBO–MJO link and the possible mechanism are also discussed. Subseasonal Forecasts of Convectively Coupled Equatorial Waves and the MJO: Activity and Predictive Skill Abstract: In this study, the contribution of low-frequency (.100 days), Madden–Julian oscillation (MJO), and con-vectively coupled equatorial wave (CCEW) variability to the skill in predicting convection and winds in the tropics at weeks 1–3 is examined. We use subseasonal forecasts from the Navy Earth System Model (NESM);NCEP Climate Forecast System, version 2 (CFSv2); and ECMWF initialized in boreal summer 1999–2015.A technique for performing wave number–frequency filtering on subseasonal forecasts is introduced and applied to these datasets. This approach is better able to isolate regional variations in MJO forecast skill than traditional global MJO indices. Biases in the mean state and in the activity of the MJO and CCEWs are smallest in the ECMWF model. The NESM overestimates cloud cover as well as MJO, equatorial Rossby, and mixed Rossby–gravity/tropical depression activity over the west Pacific. The CFSv2 underestimates convectively coupled Kelvin wave activity. The predictive skill of the models at weeks 1–3 is examined by decomposing the forecasts into wavenumber–frequency signals to determine the modes of variability that contribute to forecast skill. All three models have a similar ability to simulate low-frequency variability but large differences in MJO skill are observed. The skill of the NESM and ECMWF model in simulating MJO-related OLR signals at week 2 is greatest over two regions of high MJO activity, the equatorial Indian Ocean and Maritime Continent, and the east Pacific. The MJO in the CFSv2 is too slow and too weak, which results in lower MJO skill in these regions. The global response to tropical heating in the Madden–Julian oscillation during the northern winter 2004 paper. Abstract: A life cycle of the Madden–Julian oscillation (MJO) was constructed, based on 21 years of outgoing long-wave radiation data. Regression maps of NCEP–NCAR reanalysis data for the northern winter show statistically significant upper-tropospheric equatorial wave patterns linked to the tropical convection anomalies, and extratropical wave patterns over the North Pacific, North America, the Atlantic, the Southern Ocean and South America. To assess the cause of the circulation anomalies, a global primitive-equation model was initialized with the observed three-dimensional (3D) winter climatological mean flow and forced with a time-dependent heat source derived from the observed MJO anomalies. A model MJO cycle was constructed from the global response to the heating, and both the tropical and extratropical circulation anomalies generally matched the observations well. The equatorial wave patterns are established in a few days, while it takes approximately two weeks for the extratropical patterns to appear. The model response is robust and insensitive to realistic changes in damping and basic state. The model tropical anomalies are consistent with a forced equatorial Rossby–Kelvin wave response to the tropical MJO heating, although it is shifted westward by approximately 20◦longitude relative to observations. This may be due to a lack of damping processes (cumulus friction) in the regions of convective heating. Once this shift is accounted for, the extratropical response is consistent with theories of Rossby wave forcing and dispersion on the climatological flow, and the pattern correlation between the observed and modelled extratropical flow is up to 0.85. The observed tropical and extratropical wave patterns account for a significant fraction of the intraseasonal circulation variance, and this reproducibility as a response to tropical MJO convection has implications for global medium-range weather prediction. The MJO‐SSW Teleconnection: Interaction Between MJO‐Forced Waves and the Midlatitude Jet Abstract: The Madden Julian Oscillation (MJO) was shown to affect both present-day Sudden Stratospheric Warming (SSW) events in the Arctic, and their future frequency under global warming scenarios, with implications to the Arctic Oscillation and mid-latitude extreme weather. This work uses a dry dynamic core model to understand the dependence of SSW frequency on the amplitude and longitudinal range of the MJO, motivated by the prediction that the MJO will strengthen and broaden its longitudinal range in a warmer climate. We focus on the response of the mid-latitude jets and the corresponding generated stationary waves, that are shown to dominate the response of SSW events to MJO forcing. Momentum budget analysis of a large ensemble of spinup simulations suggests the climatological jet response is driven by the MJO-forced meridional eddy momentum transport. The results suggest that the trends in both MJO amplitude and longitudinal range are important for the prediction of the mid-latitude jet response and for the prediction of SSWs in a future climate. The Relationship between Northern Hemisphere Winter Blocking and Tropical Modes of Variability 2016 paper. Abstract: In the present study, the influence of some major tropical modes of variability on Northern Hemisphere regional blocking frequency variability during boreal winter is investigated. Reanalysis data and an ensemble experiment with the ECMWF model using relaxation toward the ERA-Interim data inside the tropics areused. The tropical modes under investigation are El Niño–Southern Oscillation (ENSO), the Madden–Julian oscillation (MJO), and the upper-tropospheric equatorial zonal-mean zonal wind [U1^50]E. An early (late) MJO phase refers to the part of the MJO cycle when enhanced (suppressed) precipitation occurs over the western Indian Ocean and suppressed (enhanced) precipitation occurs over the Maritime Continent and the western tropical Pacific. Over the North Pacific sector, it is found that enhanced (suppressed) high-latitude blocking occurs in association with El Niño (La Niña) events, late (early) MJO phases, and westerly (easterly)[U1^50]E. Over central to southern Europe and the east Atlantic, it is found that late MJO phases, as well as a suppressed MJO, are leading to enhanced blocking frequency. Furthermore, early (late) MJO phases arefollowed by blocking anomalies over the western North Atlantic region, similar to those associated with a positive (negative) North Atlantic Oscillation. Over northern Europe, the easterly (westerly) phase of[U1^50]Eis associated with enhanced (suppressed) blocking. These results are largely confirmed by both the reanalysis and the model experiment. The Role of Zonal Asymmetry in the Enhancement and Suppression of Sudden Stratospheric Warming Variability by the Madden–Julian Oscillation Abstract: Sudden stratospheric warming (SSW) events influence the Arctic Oscillation and midlatitude extreme weather. Previous work showed the Arctic stratosphere to be influenced by the Madden–Julian oscillation(MJO) and that the SSW frequency increases with an increase of the MJO amplitude, expected in a warmer climate. It is shown here that the zonal asymmetry in both the background state and forcing plays a dominant role, leading to either enhancement or suppression of SSW events by MJO-like forcing. When applying a circumglobal MJO-like forcing in a dry dynamic core model, the MJO-forced waves can change the general circulation in three ways that affect the total vertical Eliassen–Palm flux in the Arctic stratosphere. First,weakening the zonal asymmetry of the tropospheric midlatitude jet, and therefore preventing the MJO-forced waves from propagating past the jet. Second, weakening the jet amplitude, reducing the waves generated in the midlatitudes, especially stationary waves, and therefore the upward-propagating planetary waves. Third,reducing the Arctic lower-stratospheric refractory index, which prevents waves from upward propagation.These effects stabilize the Arctic vortex and lower the SSW frequency. The longitudinal range to which the MJO-like forcing is limited plays an important role as well, and the strongest SSW frequency increase is seen when the MJO is located where it is observed in current climate. The SSW suppression effects are active when the MJO-like forcing is placed at different longitudinal locations. This study suggests that future trends in both the MJO amplitude and its longitudinal extent are important for predicting the Arctic stratosphere response. The tropical Madden-Julian oscillation and the global wind oscillation Abstract: The global wind oscillation (GWO) is a subseasonal phenomenon encompassing the Madden-Julian Oscillation (MJO) and mid-latitude processes like meridional momentum transports and mountain torques. A phase space is defined for the GWO following the approach of Wheeler and Hendon (2004) for the MJO. In contrast to the oscillatory behavior of the MJO, two red noise processes define the GWO. The red noise spectra have variance at periods that bracket the 30-60 day band generally used to define the MJO. The MJO and GWO correlation accounts for 25% of their variance and cross-spectra show well-defined phase relations. However, considerable independent variance still exists in the GWO. During MJO and GWO episodes, key events in the circulation and tropical convection derived from composites can be used for monitoring and for evaluating prediction model forecasts, especially for weeks 1-3. A case study during April-May 2007 focuses on the GWO and two ~30 day duration orbits with extreme anomalies in GWO phase space. The MJO phase space projections for the same time were partially driven by mountain torques and meridional transports. The case reveals the tropical-extratropical character of subseasonal events and its role in creating slowly evolving planetary-scale circulation and tropical convection anomalies. Time-Lagged Response of the Antarctic and High-Latitude Atmosphere to Tropical MJO Convection Abstract: Intraseasonal tropical variability has important implications for themid- and high-latitude atmosphere, and inrecent studies has been shown to modulate a number of weather processes inthe Northern Hemisphere, such as snow depth, sea ice concentration, precipitation, atmospheric rivers, and air temperature. In such studies, the extratropical atmosphere has tended to respond to the tropical convection of the leading mode of intraseasonal variability, the Madden–Julian oscillation (MJO), with a time lag of approximately 7 days. However, the time lag between the MJO and the Antarctic atmosphere has been found to vary between less than 7 and greater than 20 days. This study builds on previous work by further examining the time-lagged response of SouthernHemisphere tropospheric circulation to tropical MJO forcing, with specific focus on the latitude belt associated with the Antarctic Oscillation, during the months of June (austral winter) and December (austral summer) using NCEP–DOE Reanalysis 2 data for the years 1979–2016. Principal findings indicate that the timelag with the strongest height anomalies depends on both the location of the MJO convection (e.g., the MJO phase) and the season, and thatthe lagged height anomalies in the Antarctic atmosphere are fairly consistent across different vertical levels and latitudinal bands. In addition, certain MJO phases in December displayed lagged height anomalies indicative of blocking-type atmospheric patterns, with an approximate wavenumber of 4, whereas in June most phases were as-sociated with more progressive height anomaly centers resembling a wavenumber-3-type pattern. Tropical–extratropical interactions (section from the 2011 publication Intraseasonal Variability in the Atmosphere–Ocean ClimateSystem) No abstract, but this from theclosing 'Discussion' chapter: The MJO is associated with significant signals that extend around the globe andthroughout the seasonal cycle. Organized convection in its active convective phaseresults in a planetary wave response that modulates teleconnection patterns aroundthe globe. Convection organized in the tropics associated with the MJO is alsostrongly modulated by extratropical waves, including those that previouslyoriginated within the MJO itself. Such associations suggest the possibility thatcoupling between the tropics and the extratropics might be fundamental to theevolution and structure of the MJO. The associated global signals modulateextreme weather events. Since numerical weather prediction models tend to poorlysimulate the MJO, MJO modulation of extreme events would reduce our ability topredict such events at long range by deterministic means until simulation of the MJOimproves. MJO Prediction Skill of the Subseasonal-to-Seasonal Prediction Models Published May 2018. Abstract: The Madden–Julian oscillation (MJO), the dominant mode of tropical intraseasonal variability, provides a major source of tropical and extratropical predictability on a subseasonal time scale. This study conducts a quantitative evaluation of the MJO prediction skill in state-of-the-art operational models, participating in the subseasonal-to-seasonal (S2S) prediction project. The relationship of MJO prediction skill with model biases in the mean moisture fields and in the longwave cloud–radiation feedbacks is also investigated. The S2S models exhibit MJO prediction skill out to a range of 12 to 36 days. The MJO prediction skills in the S2S models are affected by both the MJO amplitude and phase errors, with the latter becoming more important at longer forecast lead times. Consistent with previous studies, MJO events with stronger initial MJO amplitude are typically better predicted. It is found that the sensitivity to the initial MJO phase varies notably from model to model. In most models, a notable dry bias develops within a few days of forecast lead time in the deep tropics, especially across the Maritime Continent. The dry bias weakens the horizontal moisture gradient over the Indian Ocean and western Pacific, likely dampening the organization and propagation of the MJO. Most S2S models also underestimate the longwave cloud–radiation feedbacks in the tropics, which may affect the maintenance of the MJO convective envelope. The models with smaller bias in the mean horizontal moisture gradient and the longwave cloud–radiation feedbacks show higher MJO prediction skills, suggesting that improving those biases would enhance MJO prediction skill of the operational models.
    1 point
  8. Here are the current Papers & Articles under the research topic Brewer-Dobson Circulation. Click on the title of a paper you are interested in to go straight to the full paper. Papers and articles covering the basics (ideal for learning) are shown in Green. (About the) Brewer-Dobson Circulation The Brewer‐Dobson circulation What drives the Brewer-Dobson Circulation? The Brewer–Dobson Circulation: Dynamics of the Tropical Upwelling Brewer–Dobson Circulation: Recent-Past and Near-Future Trends Simulated by Chemistry-Climate Models
    1 point
This leaderboard is set to London/GMT+01:00
×
×
  • Create New...