The Weather Game

Ieuan Higgs – i.higgs@pgr.reading.ac.uk

It’s a colder-than-usual, early October Friday afternoon in the PhD offices of Brian Hoskins. The week is tired, motivation is waning and most importantly – Sappo is only 30 minutes away. As the collective mind of each office meanders further and further from work, someone inevitably pipes up with:

“Has anyone done their weather game predictions this week?”

Some mutterings might move around the room – grumbling about the unpredictability of rainfall in Singapore, or a verbal jab at the cold front that decided to move across Reading about 6 hours too early – until, as predictably as ENSO, a first year cautiously asks,

“…What’s the weather game?”

Which is then met with a suitable response, such as:

“The Weather Game? It’s a bit like fantasy football, but for us weather nerds – you’re going to love it!”. 

At least, that’s how I like to describe it.

The game was hotly contested this Autumn, with huge sign-up and participation across the entire term.

A particular shoutout to the Undergraduates, who were out in force and took 50% of the top 10 spots!

Plotting the cumulative scores for the top 32 players of the term, we are treated to a blindingly colourful cascade (thanks excel?) of points totals:

From this, it is clear that our eventual winner had led the pack for a solid five weeks by the competition end – although I’m sure they were a little nervous in those final two weeks. We can also see the dreaded “flatline” – players who clearly got off to a good start but then, for whatever reason, never submitted another prediction for the remainder of the game. Another interesting feature of these plots is the occasional downward bump – a symptom of the dreaded negative score, which were (thankfully) relatively few and far between.

The illustrious awards ceremony was held in WCD on the 8th of December. Category winners were treated to a bar of tasty chocolate, and the overall winner was gifted a delightful little ThermoPro Bluetooth Thermometer & Hygrometer. This seemed an ideal prize for students who might want check if their flat-share thermostat is being undemocratically switched on while they are out at lectures. Of course, a wooden spoon was given to the last place that played at least 8 of the 10 weeks (and if you’re having that much fun with the weather – can you really ever lose?).

With all of that said, we now put our winners’ names in lights (or on the blog) – immortalising them in the records of Weather Game glory.

Wooden Spoon: Catherine Toolan

Oil and Gas – 66.8 points – 32nd place

Best Pseudonym: Meg Stretton

The SIF Lord

External: Thomas Hall

Noctilucent – 518.6 points – 2nd place

Postgraduate: Caleb Miller

I own a sphere – 414.4 points – 8th place

Staff: Patrick McGuire

WindyCrashLandingOnYou – 432.4 points – 7th place

Overall Winner: Nathan Ng

Come Rain or (Keith) Shine – 534.3 points – 1st place

The Weather Game will be back in Spring of 2024. We are excited to run it, and hope to see many new and familiar faces (well, pseudonyms) there.

Mr Weathers

Ieuan Higgs and Nathan Edward-Inatimi

WesCon 2023: From Unexpected Radiosondes to Experimental Forecasts

Adam Gainford – a.gainford@pgr.reading.ac.uk

Summer might seem like a distant memory at this stage, with the “exact date of snow” drawing ever closer and Mariah Carey’s Christmas desires broadcasting to unsuspecting shoppers across the country. But cast your minds back four-to-six months and you may remember a warmer and generally sunnier time, filled with barbeques, bucket hats, and even the occasional Met Ball. You might also remember that, weather-wise, summer 2023 was one of the more anomalous summers we have experienced in the UK. This summer saw 11% more rainfall recorded than the 1991-2020 average, despite June being dominated by hot, dry weather. In fact, June 2023 was also the warmest June on record and yet temperatures across the summer turned out to be largely average. 

Despite being a bit of an unsettled summer, these mixed conditions provided the perfect opportunity to study a notoriously unpredictable type of weather: convection. Convection is often much more difficult to accurately forecast compared to larger-scale features, even using models which can now explicitly resolve these events. As a crude analogy, consider a pot of bubbling water which has brought to the boil on a kitchen hob. As the amount of heat being delivered to the water increases, we can probably make some reasonable estimates of the number of bubbles we should expect to see on the surface of the water (none initially, but slowly increasing in number as the temperature of the water approaches the boiling point). But we would likely struggle if we tried to predict exactly where those bubbles might appear. 

This is where the WesCon (Wessex Convection) field campaign comes in. WesCon participants spent the entire summer operating radars, launching radiosondes, monitoring weather stations, analysing forecasts, piloting drones, and even taking to the skies — all in an effort to better understand convection and its representation within forecast models. It was a huge undertaking, and I was fortunate enough to be a small part of it. 

In this blog I discuss two of the ways in which I was involved: launching radiosondes from the University of Reading Atmospheric Observatory and evaluating the performance of models at the Met Office Summer Testbed.

Radiosonde Launches and Wiggly Profiles

A core part of WesCon was frequent radiosonde launches from sites across the south and south-west of the UK. Over 300 individual sondes were launched in total, with each one requiring a team of two to three people to calibrate the sonde, record station measurements and fill balloons with helium. Those are the easy parts – the hard part is making sure your radiosonde gets off the ground in one piece.

You can see in the picture below that the observatory is surrounded by sharp fences and monitoring equipment which can be tricky to avoid, especially during gusty conditions. In the rare occurrences when the balloon experienced “rapid unplanned disassembly”, we had to scramble to prepare a new one so as not to delay the recordings by too long.

The University of Reading Atmospheric Observatory, overlooked by some mid-level cloud streets. 

After a few launches, however, the procedure becomes routine. Then you can start taking a cursory look at the data being sent back to the receiving station.

During the two weeks I was involved with launching radiosondes, there were numerous instances of elevated convection, which were a particular priority for the campaign given the headaches these cause for modellers. Elevated convection is where the ascending airmass originates from somewhere above the boundary layer, such as on a frontal boundary. We may therefore expect profiles of elevated convection to include a temperature inversion of some kind, which would prevent surface airmasses from ascending above the boundary layer. 

However, what we certainly did not expect to see were radiosondes appearing to oscillate with height (see my crude screenshot below). 

“The wiggler”! Oscillating radiosondes observed during elevated convection events.

Cue the excited discussions trying to explain what we were seeing. Sensor malfunction? Strong downdraughts? Not quite. 

Notice that the peak of each oscillation occurs almost exactly at 0°C. Surely that can’t be coincidental! Turns out these “wiggly” radiosondes have been observed before, albeit infrequently, and is attributed to snow building up on the surface of the balloon, weighing it down. As the balloon sinks and returns to above-freezing temperatures, the accumulated snow gradually melts and departs the balloon, allowing it to rise back up to the freezing level and accumulate more snow, and so on. 

That sounds reasonable enough. So why, then, do we see this oscillating behaviour so infrequently? One of the reasons discovered was purely technical. 

If you would like to read more about these events, a paper is currently being prepared by Stephen Burt, Caleb Miller and Brian Lo. Check back on the blog for further updates!

Humphrey Lean, Eme Dean-Lewis (left) and myself (right) ready to launch a sonde.

Met Office Summer Testbed

While not strictly a part of WesCon, this summer’s Met Office testbed was closely connected to the themes of the field campaign, and features plenty of collaboration. 

Testbeds are an opportunity for operational meteorologists, researchers, academics, and even students to evaluate forecast outputs and provide feedback on particular model issues. This year’s testbed was focussed on two main themes: convection and ensembles. These are both high priority areas for development in the Met Office, and the testbed provides a chance to get a broader, more subjective evaluation of these issues.

Group photo of the week 2 testbed participants.

Each day was structured into six sets of activities. Firstly, we were divided into three groups to perform a “Forecast Denial Experiment”, whereby each group is given access to a limited set of data and asked to issue a forecast for later in the day. One group only had access to the deterministic UKV model outputs, another group only had access to the MOGREPS-UK high-resolution ensemble output, and the third group has access to both datasets. The idea was to test whether ensemble outputs provide added value and accuracy to forecasts of impactful weather compared to just deterministic outputs. Each group was led by one or two operational meteorologists who navigated the data and, generally, provided most of the guidance. Personally, I found it immensely useful to shadow the op-mets as they made their forecasts, and came away with a much better understanding of the processes which goes into issuing a forecast.

After lunch, we would begin the ensemble evaluation activity which focussed on subjectively evaluating the spread of solutions in the high-resolution MOGREPS-UK ensemble. Improving ensemble spread is one of the major priorities for model development; currently, the members of high-resolution ensembles tend to diverge from the control member too slowly, leading to overconfident forecasts. It was particularly interesting to compare the spread results from MOGREPS-UK with the global MOGREPS-G ensemble and to try to understand the situations when the UK ensemble seemed to resemble a downscaled version of the global model. Next, we would evaluate three surface water flooding products, all combining ensemble data with other surface and impact libraries to produce flooding risk maps. Despite being driven by the same underlying model outputs, it was surprising how much each model differed in the case studies we looked at. 

Finally, we would end the day by evaluating the WMV (Wessex Model Variable) 300 m test ensemble, run over the greater Bristol area over this summer for research purposes. Also driven by MOGREPS-UK, this ensemble would often pick out convective structure which MOGREPS-UK was too coarse to resolve, but also tended to overdo the intensities. It was also very interesting to see the objective metrics suggested that WMV had much worse spread than MOGREPS-UK over the same area, a surprising result which didn’t align with my own interpretation of model performance.

Overall, the testbed was a great opportunity to learn more about how forecasts are issued and to get a deeper intuition for how to interpret model outputs. As researchers, it’s easy to look at model outputs as just abstract data, which is there to be verified and scrutinised, forgetting the impacts that it can have on the people experiencing it. While it was an admittedly exhausting couple of weeks, I would highly recommend more students take part in future testbeds!

Machine Learning: complement or replacement of Numerical Weather Prediction? 

Emanuele Silvio Gentile – e.gentile@pgr.reading.ac.uk

Figure 1 Replica of the first 1643 Torricelli barometer [1]

Humans have tried, for millennia, to predict the weather by finding physical relationships between observed weather events, a notable example being the descent in barometric pressure used as an indicator of an upcoming precipitation event. It should come as no surprise that one of the first weather measuring instrument to be invented was the barometer, by Torricelli (see in Fig. 1 a replica of the first Torricelli barometer), nearly concurrently with a reliable thermometer. Only two hundred years later, the development of the electric telegraph allowed for a nearly instant exchange of weather data, leading to the creation of the first synoptic weather maps in the US, followed by Europe. Synoptic maps allowed amateur and professional meteorologists to look at patterns between weather data in an unprecedented effective way for the time, allowing the American meteorologists Redfield and Epsy to resolve the dispute on which way the air flowed in a hurricane (anticlockwise in the Northern Hemisphere).

Figure 2 High Resolution NWP – model concept [2]

By the beginning of the 20th century many countries around the globe started to exchange data daily (thanks to the recently laid telegraphic cables) leading to the creation of global synoptic maps, with information in the upper atmosphere provided by radiosondes, aeroplanes, and in the 1930s radars. By then, weather forecasters had developed a large set of experimental and statistical rules on how to compute the changes to daily synoptic weather maps looking at patterns between historical sets of synoptic daily weather maps and recorded meteorological events, but often, prediction of events days in advance remained challenging.

In 1954, a powerful tool became available to humans to objectively compute changes on the synoptic map over time: Numerical Weather Prediction models. NWPs solve numerically the primitive equations, a set of nonlinear partial differential equations that approximate the global atmospheric flow, using as initial conditions a snapshot of the state of the atmosphere, termed analysis, provided by a variety of weather observations. The 1960s, marked by the launch of the first satellites, enabled 5-7 days global NWP forecasts to be performed. Thanks to the work of countless scientists over the past 40 years, global NWP models, running at a scale of about 10km, can now simulate skilfully and reliably synoptic-scale and meso-scale weather patterns, such as high-pressure systems and midlatitude cyclones, with up to 10 days of lead time [3].

The relatively recent adoption of limited-area convection-permitting models (Fig. 2) has made possible even the forecast of local details of weather events. For example, convection-permitting forecasts of midlatitude cyclones can accurately predict small-scale multiple slantwise circulations, the 3-D structure of convection lines, and the peak cyclone surface wind speed [4].

However, physical processes below convection permitting resolution, such as wind gusts, that present an environmental risk to lives and livelihoods, cannot be explicitly resolved, but can be derived from the prognostic fields such as wind speed and pressure. Alternative techniques, such as statistical modelling (Malone model), haven’t yet matched (and are nowhere near to) the power of numerical solvers of physical equations in simulating the dynamics of the atmosphere in the spatio-temporal dimension.

Figure 3 Error growth over time [5]

NWPs are not without flaws, as they are affected by numerical drawbacks: errors in the prognostic atmospheric fields build up through time, as shown in Fig. 3, reaching a comparable forecast error to that of a persisted forecast, i.e. at each time step the forecast is constant, and of a climatology-based forecast, i.e. mean based on historical series of observations/model outputs. Errors build up because NWPs iteratively solve the primitive equations approximating the atmospheric flow (either by finite differences or spectral methods). Sources of these errors are: too coarse model resolution (which leads to incorrect representation of topography), long integration time steps, and small-scale/sub-grid processes which are unresolved by the model physics approximations. Errors in parametrisations of small-scale physical processes grow over time, leading to significant deterioration of the forecast quality after 48h. Therefore, high-fidelity parametrisations of unresolved physical processes are critical for an accurate simulation of all types of weather events.

Figure 4 Met Office HPC [6]

Another limitation of NWPs is the difficulty in simulating the chaotic nature of weather, which leads to errors in model initial conditions and model physics approximations that grow exponentially over time. All these limitations, combined with instability of the atmosphere at the lower and upper bound, make the forecast of rapidly developing events such as flash floods particularly challenging to predict. A further weakness of NWP forecasts is that they rely on the use of an expensive High Parallel Computing (HPC) facility (Fig. 4), owned by a handful of industrialised nations, which run coarse scale global models and high-resolution convection-permitting forecasts on domains covering area of corresponding national interest. As a result, a high resolution prediction of weather hazards, and climatological analysis remains off-limits for the vast majority of developing and third-world countries, with detrimental effects not just on first line response to weather hazards, but also for the development of economic activities such agriculture, fishing, and renewable energies in a warming climate. In the last decade, the cloud computing technological revolution led to a tremendous increase in the availability and shareability of weather data sets, which transitioned from local storage and processing to network-based services managed by large cloud computing companies, such as Amazon, Microsoft or Google, through their distributed infrastructure.

Combined with the wide availability of their cloud computing facilities, the access to weather data has become more and more democratic and ubiquitous, and consistently less dependent on HPC facilities owned by National Agencies. This transformation is not without drawbacks in case these tech giants decide to close the taps of the flow of data. During a row with the Australian government, Facebook banned access to Australian news content in Feb 2021. Although by accident, also government-related agencies such as the Bureau of Meteorology were banned, leaving citizens with restricted access to important weather information until the pages were restored. It is hoped that with more companies providing distributed infrastructure, the accessibility to vital data for citizen security will become more resilient.

The exponential accessibility of weather data sets has stimulated the development and the application of novel machine learning algorithms. As a result, weather scientists worldwide can crunch increasingly effectively multi-dimensional weather data, ultimately providing a new powerful paradigm to understand and predict the atmospheric flow based on finding relationships between the available large-scale weather datasets.

Machine learning (ML) finds meaningful representations of the patterns between the data through a series of nonlinear transformations of the input data. ML pattern recognition is distinguished into two types: supervised and unsupervised learning.

Figure5 Feed-forward neural network [6]

Supervised Learning is concerned with predicting an output for a given input. It is based on learning the relationship between inputs and outputs, using training data consisting in example input/output pairs, being divided into regression or classification, depending on the type of the output variable to be predicted (discrete or continuous). Support Vector Machine (SVM) or Regression (SVR), Artificial Neural Network (ANN, with the feed-forward step shown in Fig. 5), and Convolutional Neural Network (CNN) are examples of supervised learning.

Unsupervised learning is the task of finding patterns within data without the presence of any ground truth or labelling of the data, with a common unsupervised learning task being clustering (group of data points that are close to one another, relative to data points outside the cluster). Examples of unsupervised learning are the K-means and K-Nearest Neighbour (KNN) algorithms [7].

So far, ML algorithms have been applied to four key problems in weather prediction:  

  1. Correction of systematic error in NWP outputs, which involves post-processing data to remove biases [8]
  1. Assessment of the predictability of NWP outputs, evaluating the uncertainty and confidence scores of ensemble forecasting [9]
  1. Extreme detection, involving prediction of severe weather such as hail, gust or cyclones [10]
  1. NWP parametrizations, replacing empirical models for radiative transfer or boundary-layer turbulence with ML techniques [11]

The first key problem, which concerns the correction of systematic error in NWPs, is the most popular area of application of ML methods in meteorology. In this field, wind speed and precipitation observational data are often used to perform an ML linear regression on the NWP data with the end goal of enhancing its accuracy and resolving local details of the weather which were unresolved by NWP forecasts. Although attractive for its simplicity and robustness, linear regression presents two problems: (1) least-square methods used to solve linear regression do not scale well with the size of datasets (since matrix inversion required by least square is increasingly expensive for increasing datasets size), (2) Many relationships between variables of interest are nonlinear. Instead, classification tree-based methods have proven very useful to model non-linear weather events, from thunderstorm and turbulence detection to extreme precipitation events, and the representation of the circular nature of the wind. In fact, compared to linear regression, random trees exhibit an easy scalability with large-size datasets which have several input variables. Besides preserving the scalability to large datasets of tree-based method, ML methods such as ANN and SVM/R provide also a more generic and more powerful alternative for nonlinear-processes modelling. These improvements have come at the cost of a difficult interpretation of the underlying physical concepts that the model can identify, which is critical given that scientists need to couple these ML models with physical-equations based NWP for variable interdependence. As a matter of fact, it has proven challenging to interpret the physical meaning of the weights and nonlinear activation functions that describe in the ANN model the data patterns and relationships found by the model [12].

The second key problem, represented by the interpretation of ensemble forecasts, is being addressed by ML unsupervised learning methods such as clustering, which represents the likelihood of a forecast aggregating ensemble members by similarity. Examples include grouping of daily weather phenomena into synoptic types, defining weather regimes from upper air flow patterns, and grouping members of forecast ensembles [13].

The third key problem, which concerns the prediction of weather extremes, corresponding to weather phenomena which are a hazard to lives and economic activities, ML based methods tend to underestimate these events. The problem here lies with imbalanced datasets, since extreme events represent only a very small fraction of the total events observed [14].

The fourth key problem to which ML is currently being applied, is parametrisation. Completely new stochastic ML approaches have been developed, and their effectiveness, along with their simplicity compared to traditional empirical models has highlighted promising future applications in (moist) convection [15]

Further applications of ML methods are currently limited by intrinsic problems affecting the ML methods in relation to the challenges posed by weather data sets. While the reduction of the dimensionality of the data by ML techniques has proven highly beneficial for image pattern recognition in the context of weather data, it leads to a marked simplification of the input weather data, since it constrains the input space to individual grid cells in space or time [16]. The recent expansion of ANN into deep learning has provided new methodologies that can address these issues. This has pushed further the capability of ML models within the weather forecast domain, with CNNs providing a methodology for extracting complex patterns from large, structured datasets have been proposed, an example being the CNN model developed by Yunjie Liu in 2016 [17] to classify atmospheric rivers from climate datasets (atmospheric rivers are an important physical process for prediction of extreme rainfall events).

Figure 7 Sample images of atmospheric rivers correctly classified (true positive) by the deep CNN model in [18]

At the same time, Recursive Neural Networks (RNN), developed for natural language processing, are improving nowcasting techniques exploiting their excellent ability to work with the temporal dimension of data frames. CNN and RNN have now been combined, as illustrated in Fig. 6, providing the first nowcasting method in the context of precipitation, using radar data frames as input [18].

Figure 6 Encoding-forecasting ConvLSTM network for precipitation nowcasting [18]

While these results show a promising application of ML models to a variety of weather prediction tasks which extend beyond the area of competence of traditional NWPs, such as analysis of ensemble clustering, bias correction, analysis of climate data sets and nowcasting, they also show that ML models are not ready to replace NWP to forecast synoptic-scale and mesoscale weather patterns.

As a matter of fact, NWPs have been developed and improved for over 60 years with the very purpose to simulate very accurately and reliably the wind, pressure, temperature and other relevant prognostic fields, so it would be unreasonable to expect ML models to outperform NWPs on such tasks.

It is also true that, as noted earlier, the amount of available data will only grow in the coming decades, so it will be critical as well as strategic to develop ML models capable to extract patterns and interpret the relationships within such data sets, complementing NWP capabilities. But how long before an ML model will be capable to replace an NWP by crunching the entire set of historical observations of the atmosphere, extracting the patterns and the spatial-temporal relationships between the variables, and then performing weather forecasts?

Acknowledgement: I would like to thank my colleagues and friends Brian Lo, James Fallon, and Gabriel M. P. Perez, for reading and providing feedback on this article.

References

  1. https://collection.sciencemuseumgroup.org.uk/objects/co54518/replica-of-torricellis-first-barometer-1643-barometer-replica 
  1. https://www.semanticscholar.org/paper/High-resolution-numerical-weather-prediction-(NWP)-Allan-Bryan/a40e0ebd388b915bdd357f398baa813b55cef727/figure/6 
  1. Buizza, R., Houtekamer, P., Pellerin, G., Toth, Z., Zhu, Y. and Wei, M. (2005) A comparison of the ECMWF, MSC, and NCEP global ensemble prediction systems. Mon Weather Rev, 133, 1076 – 1097 
  1. Lean, H. and Clark, P. (2003) The effects of changing resolution on mesocale modelling of line convection and slantwise circulations in FASTEX IOP16. Q J R Meteorol Soc, 129, 2255–2278 
  1. http://www.chanthaburi.buu.ac.th/~wirote/met/tropical/textbook_2nd_edition/navmenu.php_tab_10_page_4.3.5.htm 
  1. Bishop, C., and Christopher, M., Pattern Recognition and Machine Learning, Springer 
  1. https://www.arup.com/projects/met-office-high-performance-computer 
  1. J. L. Aznarte and N. Siebert, “Dynamic Line Rating Using Numerical Weather Predictions and Machine Learning: A Case Study,” in IEEE Transactions on Power Delivery, vol. 32, no. 1, pp. 335-343, Feb. 2017, doi: 10.1109/TPWRD.2016.2543818. 
  1. Foley, Aoife M et al. (2012). “Current methods and advances in forecasting of wind power generation”. In: Renewable Energy 37.1, pp. 1–8. 
  1. McGovern, Amy et al. (2017). “Using artificial intelligence to improve real-time decision making for high-impact weather”. In: Bulletin of the American Meteorological Society 98.10, pp. 2073–2090 
  1. O’Gorman, Paul A and John G Dwyer (2018). “Using machine learning to parameterize moist convection: Potential for modeling of climate, climate change and extreme events”. In: arXiv preprint arXiv:1806.11037 
  1. Moghim, Sanaz and Rafael L Bras (2017). “Bias correction of climate modeled temperature and precipitation using artificial neural networks”. In: Journal of Hydrometeorology 18.7, pp. 1867–1884.  
  1. Camargo S J, Robertson A W Gaffney S J Smyth P and M Ghil (2007). “Cluster analysis of typhoon tracks. Part I: General properties”. In: Journal of Climate 20.14, pp. 3635–3653. 
  1. Ahijevych, David et al. (2009). “Application of spatial verification methods to idealized and NWP-gridded precipitation forecasts”. In: Weather and Forecasting 24.6, pp. 1485–1497. 
  1. Berner, Judith et al. (2017). “Stochastic parameterization: Toward a new view of weather and climate models”. In: Bulletin of the American Meteorological Society 98.3, pp. 565–588. 
  1. Fan, Wei and Albert Bifet (2013). “Mining big data: current status, and forecast to the future”. In: ACM sIGKDD Explorations Newsletter 14.2, pp. 1–5 
  1. Liu, Yunjie et al. (2016). “Application of deep convolutional neural networks for detecting extreme weather in climate datasets”. In: arXiv preprint arXiv:1605.01156. 
  1. Xingjian, SHI et al. (2015). “Convolutional LSTM network: A machine learning approach for precipitation nowcasting”. In: Advances in neural information processing systems, pp. 802–810. 

Diagnosing solar wind forecast errors

Harriet Turner – h.turner3@pgr.reading.ac.uk

The solar wind is a continual outflow of charged particles that comes off the Sun, ranging in speed from 250 to 800 km s-1. During the first six months of my PhD, I have been investigating the errors in a type of solar wind forecast that uses spacecraft observations, known as corotation forecasts. This was the topic of my first paper, where I focussed on extracting the forecast error that occurs due to a separation in the spacecraft latitude. I found that up to a latitudinal separation of 6 degrees, the error contribution was approximately constant. Above 6 degrees, the error contribution increases as the latitudinal separation increases. In this blog post I will explain the importance of forecasting the solar wind and the principle behind corotation forecasts. I will also explain how this work has wider implications for future space missions and solar wind forecasting.

The term “space weather” refers to the changing conditions in near-Earth space. Extreme space weather events can cause several effects on Earth, such as damaging power grids, disrupting communications, knocking out satellites and harming the health of humans in space or on high-altitude flights (Cannon, 2013). These effects are summarised in Figure 1. It is therefore important to accurately forecast space weather to help mitigate against these effects. Knowledge of the background solar wind is an important aspect of space weather forecasting as it modulates the severity of extreme events. This can be achieved through three-dimensional computer simulations or through more simple methods, such as corotation forecasts as discussed below.

Figure 1. Cosmic rays, solar energetic particles, solar flare radiation, coronal mass ejections and energetic radiation belt particles cause space weather. Subsequently, this produces a number of effects on Earth. Source: ESA.

Solar wind flow is mostly radial away from the Sun, however the fast/slow structure of the solar wind rotates round with the Sun. If you were looking down on the ecliptic plane (where the planets lie, at roughly the Sun’s equator), then you would see a spiral shape of fast and slow solar wind, as in Figure 2. This makes a full rotation in approximately 27 days. As this rotates around, it allows us to use observations on this plane as a forecast for a point further on in that rotation, assuming a steady-state solar wind (i.e., the solar wind does not evolve in time). For example, in Figure 2, an observation from the spacecraft represented by the red square could be used as a forecast at Earth (blue circle), some time later. This time depends on the longitudinal separation between the two points, as this determines the time it takes for the Sun to rotate through that angle.

Figure 2. The spiral structure of the solar wind, which rotates anticlockwise. Here, STA and STB are the STEREO-A and STEREO-B spacecraft respectively. The solar wind shown here is the radial component. Source: HUXt model (Owens et al, 2020).

In my recent paper I have been investigating how the corotation forecast error varies with the latitudinal separation of the observation and forecast points.  Latitudinal separation varies throughout the year, and it was theorised that it should have a significant impact on the accuracy of corotation forecasts. I used the two spacecraft from the STEREO mission, which are on the same plane as Earth, and a dataset for near-Earth. This allowed for six different configurations to compute corotation forecasts, with a maximum latitudinal separation of 14 degrees. I analysed the 18-month period from August 2009 to February 2011 to help eliminate other affecting variables. Figure 3 shows the relationship between forecast error and latitudinal separation. Up to approximately 6 degrees, there is no significant relationship between error and latitudinal separation. Above this, however, the error increases approximately linearly with the latitudinal separation.

Figure 3. Variation of forecast error with the latitudinal separation between the spacecraft making the observation and the forecast location. Error bars span one standard error on the mean.

This work has implications for the future Lagrange space weather monitoring mission, due for launch in 2027. The Lagrange spacecraft will be stationed in a gravitational null, 60degrees in longitude behind Earth on the ecliptic plane. Gravitational nulls occur when the gravitational fields between two or more massive bodies balance out. There are five of these nulls, called the Lagrange points, and locating a spacecraft at one reduces the amount of fuel needed to stay in position. The goal of the Lagrange mission is to provide a side-on view of the Sun-Earth line, but it also presents an opportunity for consistent corotation forecasts to be generated at Earth. However, the Lagrange spacecraft will oscillate in latitude compared to Earth, up to a maximum of about 5 degrees. My results indicate that the error contribution from latitudinal separation would be approximately constant.

The next steps are to use this information to help improve the performance of solar wind data assimilation. Data assimilation (DA) has led to large improvements in terrestrial weather forecasting and is beginning to be used in space weather forecasting. DA combines observations and model output to find an optimum estimation of reality. The latitudinal information found here can be used to inform the DA scheme how to better handle the observations and to, hopefully, produce an improved solar wind representation.

The work I have discussed here has been accepted into the AGU Space Weather journal and is available at https://agupubs.onlinelibrary.wiley.com/doi/epdf/10.1029/2021SW002802.

References

Cannon, P.S., 2013. Extreme space weather – A report published by the UK royal academy of engineering. Space Weather, 11(4), 138-139.  https://agupubs.onlinelibrary.wiley.com/doi/full/10.1002/swe.20032

ESA, 2018. https://www.esa.int/ESA_Multimedia/Images/2018/01/Space_weather_effects 

Owens, M.J., Lang, M.S., Barnard, L., Riley, P., Ben-Nun, M., Scott, C.J., Lockwood, M., Reiss, M.A., Arge, C.N. & Gonzi, S., 2020. A Computationally Efficient, Time-Dependent Model of the Solar Wind for use as a Surrogate to Three-Dimensional Numerical Magnetohydrodynamic Simulations. Solar Physics, 295(3), https://doi.org/10.1007/s11207-020-01605-3

Connecting Global to Local Hydrological Modelling Forecasting – Virtual Workshop

Gwyneth Matthews g.r.matthews@pgr.reading.ac.uk
Helen Hooker h.hooker@pgr.reading.ac.uk 

ECMWF- CEMS – C3S – HEPEX – GFP 

What was it? 

The workshop was organised under the umbrella of ECMWF, the Copernicus services CEMS and C3S, the Hydrological Ensemble Prediction EXperiment (HEPEX) and the Global Flood Partnership (GFP). The workshop lasted 3 days, with a keynote speaker followed by Q&A at the start of each of the 6 sessions. Each keynote talk focused on a different part of the forecast chain, from hybrid hydrological forecasting to the use of forecasts for anticipatory humanitarian action, and how the global and local hydrological scales could be linked. Following this were speedy poster pitches from around the world and poster presentations and discussion in the virtual ECMWF (Gather.town).  

Figure 1: Gather.town was used for the poster sessions and was set up to look like the ECMWF site in Reading, complete with a Weather Room and rubber ducks. 

What was your poster about? 

Gwyneth – I presented Evaluating the post-processing of the European Flood Awareness System’s medium-range streamflow forecasts in Session 2 – Catchment-scale hydrometeorological forecasting: from short-range to medium-range. My poster showed the results of the recent evaluation of the post-processing method used in the European Flood Awareness System. Post-processing is used to correct errors and account for uncertainties in the forecasts and is a vital component of a flood forecasting system. By comparing the post-processed forecasts with observations, I was able to identify where the forecasts were most improved.  

Helen – I presented An evaluation of ensemble forecast flood map spatial skill in Session 3 – Monitoring, modelling and forecasting for flood risk, flash floods, inundation and impact assessments. The ensemble approach to forecasting flooding extent and depth is ideal due to the highly uncertain nature of extreme flooding events. The flood maps are linked directly to probabilistic population impacts to enable timely, targeted release of funding. The Flood Foresight System forecast flood inundation maps are evaluated by comparison with satellite based SAR-derived flood maps so that the spatial skill of the ensemble can be determined.  

Figure 2: Gwyneth (left) and Helen (right) presenting their posters shown below in the 2-minute pitches. 

What did you find most interesting at the workshop? 

Gwyneth – All the posters! Every session had a wide range of topics being presented and I really enjoyed talking to people about their work. The keynote talks at the beginning of each session were really interesting and thought-provoking. I especially liked the talk by Dr Wendy Parker about a fitness-for-purpose approach to evaluation which incorporates how the forecasts are used and who is using the forecast into the evaluation.  

Helen – Lots! All of the keynote talks were excellent and inspiring. The latest developments in detecting flooding from satellites include processing the data using machine learning algorithms directly onboard, before beaming the flood map back to earth! If openly available and accessible (this came up quite a bit) this will potentially rapidly decrease the time it takes for flood maps to reach both flood risk managers dealing with the incident and for use in improving flood forecasting models. 

How was your virtual poster presentation/discussion session? 

Gwyneth – It was nerve-racking to give the mini-pitch to +200 people, but the poster session in Gather.town was great! The questions and comments I got were helpful, but it was nice to have conversations on non-research-based topics and to meet some of the EC-HEPEXers (early career members of the Hydrological Ensemble Prediction Experiment). The sessions felt more natural than a lot of the virtual conferences I have been to.  

Helen – I really enjoyed choosing my hairdo and outfit for my mini self. I’ve not actually experienced a ‘real’ conference/workshop but compared to other virtual events this felt quite realistic. I really enjoyed the Gather.town setting, especially the duck pond (although the ducks couldn’t swim or quack! J). It was great to have the chance talk about my work and meet a few people, some thought-provoking questions are always useful.  

Forecasting space weather using “similar day” approach

Carl Haines – carl.haines@pgr.reading.ac.uk

Space weather is a natural threat that requires good quality forecasting with as much lead time as possible. In this post I outline the simple and understandable analogue ensemble (AnEn) or “similar day” approach to forecasting. I focus mainly on exploring the method itself and, although this work forecasts space weather through a timeseries of ground level observations, AnEn can be applied to many prediction tasks, particularly time series with strong auto-correlation. AnEn has previously been used to predict wind speed [1], temperature [1] and solar wind [2]. The code for AnEn is available at https://github.com/Carl-Haines/AnalogueEnsemble should you wish to try out the method for you own application. 

The idea behind AnEn is to take a set of recent observations, look back in a historic dataset for analogous periods, then take what happened following those analogous periods as the forecast. If multiple analogous periods are used, then an ensemble of forecasts can be created giving a distribution of possible outcomes with probabilistic information. 

Figure 1 – An example of AnEn applied to a space weather event with forecast time t0. The black line shows the observations, the grey line shows the ensemble members, the red line shows the median of the ensemble and the yellow and green lines are reference forecasts. 

Figure 1 is an example of a forecast made using the AnEn method where the forecast is made at t0. The 24-hours of observations (black) prior to tare matched to similar periods in the historic dataset (grey). Here I have chosen to give the most recent observations the most weighting as they hold the most relevant information. The grey analogue lines then flow on after t0 forming the forecast. Combined, these form an ensemble and the median of these is shown in red. The forecast can be chosen to be the median (or any percentile) of the ensemble or a probability of an event occurring can be given by counting how many of the ensemble member do/don’t experience the event.  

Figure 1 also shows two reference forecasts, namely 27-day recurrence and climatology, as benchmarks to beat. 27-day recurrence uses the observation from 27-days ago as the forecast for today. This is reasonable because the Sun rotates every 27-days as seen from earth so broadly speaking the same part of the Sun is emitting the relevant solar wind on timescales larger than 27-days. 

To quantify how well AnEn works as a forecast I ran the forecast on the entire dataset by repeatedly changing the forecast time t0 and applied two metrics, namely mean absolute error (MAE) and skill, to the median of the ensemble members. MAE is the size of the mean difference between the forecast made by AnEn and what was actually observed. The mean of the absolute errors over all the forecasts (taken as median of the ensemble) is taken and we end up with a value for each lead time. Figure 2 shows the MAE for AnEn median and the reference forecasts. We see that AnEn has the smallest (best) MAE at short lead times and outperforms the reference forecasts for all lead times up to a week. 

Figure 2 – The mean absolute error of the AnEn median and reference forecasts.

An error metric such as MAE cannot take into account that certain conditions are inherently more difficult to forecast such as storm times. For this we can use a skill metric defined by  

{\text{Skill} = 1 - \frac{\text{Forecast error}}{\text{Reference error}}}

where in this case we use climatology as the reference forecast. Skill can take any value between -\infty and 1 where a perfect forecast would receive a value of 1 and an unskilful forecast would receive a value of 0. A negative value of skill signifies that the forecast is worse than the reference forecast. 

Figure 3 shows the skill of AnEn and 27-day recurrence with respect to climatology. We see that AnEn is most skilful for short lead times and outperforms 27-day recurrence for all lead times considered.  

Figure 3 – The skill of the AnEn median and 27-day recurrence with respect to climatology.

In summary, the analogue ensemble forecast method matches current conditions with historical events and lifts the previously seen timeseries as the prediction. AnEn seems to perform well for this application and outperforms the reference forecasts of climatology and 27-day recurrence. The code for AnEn is available at https://github.com/Carl-Haines/AnalogueEnsemble

The work presented here makes up a part of a paper that is under review in the journal of Space Weather. 

Here, AnEn has been applied to a dataset from the space weather domain. If you would like to find out more about space weather then take a look at these previous blog posts from Shannon Jones (https://socialmetwork.blog/2018/04/13/the-solar-stormwatch-citizen-science-project/) and I (https://socialmetwork.blog/2019/11/15/the-variation-of-geomagnetic-storm-duration-with-intensity/). 

[1] Delle Monache, L., Eckel, F. A., Rife, D. L., Nagarajan, B., & Searight, K.(2013) Probabilistic Weather Prediction with an Analog Ensemble. doi: 10.1175/mwr-d-12-00281.1 

[2] Owens, M. J., Riley, P., & Horbury, T. S. (2017a). Probabilistic Solar Wind and Ge-704omagnetic Forecasting Using an Analogue Ensemble or “Similar Day” Approach. doi: 10.1007/s11207-017-1090-7 

Extending the predictability of flood hazard at the global scale

Email: rebecca.emerton@reading.ac.uk

When I started my PhD, there were no global scale operational seasonal forecasts of river flow or flood hazard. Global overviews of upcoming flood events are key for organisations working at the global scale, from water resources management to humanitarian aid, and for regions where no other local or national forecasts are available. While GloFAS (the Global Flood Awareness System, run by the European Centre for Medium-Range Weather Forecasts (ECMWF) and the European Commission Joint Research Centre (JRC) as part of the Copernicus Emergency Management Services) was producing operational, openly-available flood forecasts out to 30 days ahead, there was a need for more extended-range forecast information. Often, due to a lack of hydrological forecasts, seasonal rainfall forecasts are used as a proxy for flood hazard – however, the link between precipitation and floodiness is nonlinear, and recent research has shown that seasonal rainfall forecasts are not necessarily the best indicator of potential flood hazard. The aim of my PhD research was to look into ways in which we could provide earlier warning information, several weeks to months ahead, using hydrological analysis in addition to the meteorology.

Presidente Kuczynski recorre zonas afectadas por lluvias e inund
Flooding in Trujillo, Peru, March 2017 (Photo: Presidencia Perú on Twitter)

Broadly speaking, there are two key ways in which to provide early warning information on seasonal timescales: (1) through statistical analysis based on large-scale climate variability and teleconnections, and (2) by producing dynamical seasonal forecasts using coupled ocean-atmosphere GCMs. Over the past 4.5 years, I worked on providing hydrologically-relevant seasonal forecast products using these two approaches, at the global scale. This blog post will give a quick overview of the two new forecast products we produced as part of this research!

Can we use El Niño to predict flood hazard?

ENSO (the El Niño Southern Oscillation), is known to influence river flow and flooding across much of the globe, and often, statistical historical probabilities of extreme precipitation during El Niño and La Niña (the extremes of ENSO climate variability) are used to provide information on likely flood impacts. Due to its global influence on weather and climate, we decided to assess whether it is possible to use ENSO as a predictor of flood hazard at the global scale, by assessing the links between ENSO and river flow globally, and estimating the equivalent historical probabilities for high and low river flow, to those that are already used for meteorological variables.

With a lack of sufficient river flow observations across much of the globe, we needed to use a reanalysis dataset – but global reanalysis datasets for river flow are few and far between, and none extended beyond ~40 years (which includes a sample of ≤10 El Niños and ≤13 La Niñas). We ended up producing a 20th Century global river flow reconstruction, by forcing the Camaflood hydrological model with ECMWF’s ERA-20CM atmospheric reconstruction, to produce a 10-member river flow dataset covering 1901-2010, which we called ERA-20CM-R.

elnino_flood_hazard_gif_beccalize

Using this dataset, we calculated the percentage of past El Niño and La Niña events, during which the monthly mean river flow exceeded a high flow threshold (the 75th percentile of the 110-year climatology) or fell below a low flow threshold (the 25th percentile), for each month of an El Niño / La Niña. This percentage is then taken as the probability that high or low flow will be observed in future El Niño/La Niña events. Maps of these probabilities are shown above, for El Niño, and all maps for both El Niño and La Niña can be found here. When comparing to the same historical probabilities calculated for precipitation, it is evident that additional information can be gained from considering the hydrology. For example, the River Nile in northern Africa is likely to see low river flow, even though the surrounding area is likely to see more precipitation – because it is influenced more by changes in precipitation upstream. In places that are likely to see more precipitation but in the form of snow, there would be no influence on river flow or flood hazard during the time when more precipitation is expected. However, several months later, there may be no additional precipitation expected, but there may be increased flood hazard due to the melting of more snow than normal – so we’re able to see a lagged influence of ENSO on river flow in some regions.

While there are locations where these probabilities are high and can provide a useful forecast of hydrological extremes, across much of the globe, the probabilities are lower and much more uncertain (see here for more info on uncertainty in these forecasts) than might be useful for decision-making purposes.

Providing openly-available seasonal river flow forecasts, globally

For the next ‘chapter’ of my PhD, we looked into the feasibility of providing seasonal forecasts of river flow at the global scale. Providing global-scale flood forecasts in the medium-range has only become possible in recent years, and extended-range flood forecasting was highlighted as a grand challenge and likely future development in hydro-meteorological forecasting.

To do this, I worked with Ervin Zsoter at ECMWF, to drive the GloFAS hydrological model (Lisflood) with reforecasts from ECMWF’s latest seasonal forecasting system, SEAS5, to produce seasonal forecasts of river flow. We also forced Lisflood with the new ERA5 reanalysis, to produce an ERA5-R river flow reanalysis with which to initialise Lisflood, and to provide a climatology. The system set-up is shown in the flowchart below.

glofas_seasonal_flowchart_POSTER_EGU

I also worked with colleagues at ECMWF to design forecast products for a GloFAS seasonal outlook, based on a combination of features from the GloFAS flood forecasts, and the EFAS (the European Flood Awareness System) seasonal outlook, and incorporating feedback from users of EFAS.

After ~1 year of working on getting the system set up and finalising the forecast products, including a four-month research placement at ECMWF, the first GloFAS -Seasonal forecast was released in November 2017, with the release of SEAS5. GloFAS-Seasonal is now running operationally at ECMWF, providing forecasts of high and low weekly-averaged river flow for the global river network, up to 4 months ahead, with 3 new forecast layers available through the GloFAS interface. These provide a forecast overview for 307 major river basins, a map of the forecast for the entire river network at the sub-basin scale, and ensemble hydrographs at thousands of locations across the globe (which change with each forecast depending on forecast probabilities). New forecasts are produced once per month, and released on the 10th of each month. You can find more information on each of the different forecast layers and the system set-up here, and you can access the (openly available) forecasts here. ERA5-R, ERA-20CM-R and the GloFAS-Seasonal reforecasts are also all freely available – just get in touch! GloFAS-Seasonal will continue to be developed by ECMWF and the JRC, and has already been updated to v2.0, including a calibrated version of the hydrological model.

NEW_WEB_figure1_basins
Screenshot of the GloFAS seasonal outlook at www.globalfloods.eu

So, over the course of my PhD, we developed two new seasonal forecasts for hydrological extremes, at the global scale. You may be wondering whether they’re skilful, or in fact, which one provides the most useful forecasts! For information on the skill or ‘potential usefulness’ of GloFAS-Seasonal, head to our paper, and stay tuned for a paper coming soon (hopefully! [update: this paper has just been accepted and can be accessed online here]) on the ‘most useful approach for forecasting hydrological extremes during El Niño’, in which we compare the skill of the two forecasts at predicting observed high and low flow events during El Niño.

 

With thanks to my PhD supervisors & co-authors:

Hannah Cloke1, Liz Stephens1, Florian Pappenberger2, Steve Woolnough1, Ervin Zsoter2, Peter Salamon3, Louise Arnal1,2, Christel Prudhomme2, Davide Muraro3

1University of Reading, 2ECMWF, 3European Commission Joint Research Centre

The Circumglobal Teleconnection and its Links to Seasonal Forecast Skill for the European Summer

Email: j.beverley@pgr.reading.ac.uk

Recent extreme weather events such as the central European heatwave in 2003, flooding in the UK in 2007, and even the recent dry summer in the UK in 2018, have highlighted the need for more accurate long-range forecasts for the European summer. Recent research has led to improvements in European winter seasonal forecasts, however summer forecast skill remains relatively low. One potential source of predictability for Europe is the Indian summer monsoon, which can affect European weather via a global wave train known as the “Circumglobal Teleconnection” (CGT).

figure1
Figure 1: One-point correlation between 200 hPa geopotential height at the base point (35°-40°N, 60°-70°E) and 200 hPa geopotential height elsewhere in the ERA-Interim (1981–2014) reanalysis dataset, for August. The boxes indicate the regions defined as the “centres of action” of the CGT – these are North Pacific (NPAC), North America (NAM), Northwest Europe (NWEUR), Ding and Wang (D&W) and East Asia (EASIA).

The CGT was first identified by Ding and Wang (2005) as having a major role in modulating observed weather patterns in the Northern Hemisphere summer. Using a 200 hPa geopotential height index centred in west-central Asia (35°-40°N, 60°-70°E), they constructed a one-point correlation map of geopotential height with reference to this index (reproduced in Figure 1). From this, they identified a wavenumber-5 structure where the pressure variations over the Northeast Atlantic, East Asia, North Pacific and North America are all nearly in phase with the variations over west-central Asia (these are known as the “centres of action”). They also showed that the CGT is associated with significant temperature and precipitation anomalies in Europe, so accurate representation this mechanism in seasonal forecast models could provide an important source of subseasonal to seasonal forecast skill.

The model used here is a version of the European Centre for Medium-Range Weather Forecasts (ECMWF)’s coupled seasonal forecast model. Reforecasts are initialised on 1st May and are run for four months, so cover May-August, with start dates from 1981-2014. The skill of the model 200 hPa geopotential height is shown in Figure 2, defined as the correlation between the model ensemble mean and ERA-Interim. The model has good skill in May (to be expected given that the reforecasts are initialised in May) but in June, July and August areas of zero or negative correlation develop across much of the northern hemisphere extratropics. The areas of reduced skill align closely with the location of the centres of action of the CGT shown in Figure 1, suggesting that there is a link between the model skill and the model representation of the CGT.

figure2
Figure 2: Model ensemble mean skill for 200 hPa geopotential height as defined as the correlation between ERA-Interim and model ensemble mean for (a) May (b) June (c) July and (d) August

To determine how well the model represents the CGT, Figure 3 shows the correlation between the D&W region and the other centres of action of the CGT, as defined in Figure 1. Focussing on August (as August has the strongest CGT pattern) it can be seen that the model correlations, indicated by the box and whisker plots, are weaker than in observations (red diamond) for the D&W vs. North Pacific (NPAC), North America (NAM) and Northwest Europe (NWEUR) regions. This indicates that the model has a weak representation of the wavetrain associated with the CGT.

figure3
Figure 3: Distribution of correlation coefficients for the D&W Index correlated against the other centres of action of the CGT. The box plots represent the upper and lower quartiles, and the whiskers extend to the 5th and 95th percentiles. The black horizontal line represents the median value and the red diamond the observed correlation coefficient from ERA-Interim.

There are likely to be several reasons for the weak representation of the CGT in the model. One important factor is the presence of a northerly jet bias in the model across much of the Northern Hemisphere. This can be seen in Figure 4, which shows the model jet biases relative to ERA-Interim in the coloured contours, and the observed zonal wind in the black contours. The dipole structure of the biases which exists across much of the hemisphere, particularly in June, July and August, indicates that the model jet stream is located too far to the north. This means that Rossby waves forced in this region will have different wave propagation characteristics to reality – they may propagate at the incorrect speed, in the wrong direction or may not propagate at all, and this is likely to be an important factor in the weak representation of the CGT in the model.

figure4
Figure 4: Model 200 hPa zonal wind bias (filled contours, m/s), defined as the model ensemble mean minus ERA-Interim zonal wind, and ERA-I 200 hPa zonal wind (black contours) for (a) May (b) June (c) July and (d) August. The location of the centres of action of the CGT are marked with white crosses.

Other potential factors involved are a poor representation of the link between monsoon precipitation and the geopotential height in west-central Asia (which was shown by Ding and Wang (2007) to be important in the maintenance of the CGT) and errors in the forcing of Rossby waves associated with the monsoon. For a more detailed explanation of these, see my paper in Climate Dynamics (Beverley et al. 2018). It seems likely that the pattern of reduced skill in Figure 2, with negative correlations located at the centres of action of the CGT, including over Europe, is related to the poor representation of the CGT in the model. This raises the question of whether an improvement in the model’s representation of the CGT would lead to an improvement in forecast skill for the European summer. To address this question, sensitivity experiments have been carried out, in which the observed circulation is imposed in several centres of action along the CGT pathway to explore the impact on forecast skill for European summer weather.

References

Beverley, J. D., S. J. Woolnough, L. H. Baker, S. J. Johnson and A. Weisheimer, 2018: The northern hemisphere circumglobal teleconnection in a seasonal forecast model and its relationship to European summer forecast skill. Clim. Dyn. https://doi.org/10.1007/s00382-018-4371-4

Ding, Q., and B. Wang, 2005: Circumglobal teleconnection in the northern hemisphere summer. J. Clim. 18, 3483–3505.  https://doi.org/10.1175/JCLI3473.1

Ding, Q., and B. Wang, 2007: Intraseasonal teleconnection between the summer Eurasian wave train and the Indian monsoon. J. Clim. 20, 3751-3767. https://doi.org/10.1175/JCLI4221.1