Modelling city structure for improved urban representations in weather and climate models

By: Meg Stretton

Urban areas are home to an increasingly large proportion of the world’s population, with more people living in cities than rural areas since 2007. These large population densities mean more people are vulnerable to extreme weather events, including heatwaves, which may become more common with climate change (UK and Global extreme events – Heatwaves – Met Office). 

Extreme heatwaves may be worsened by air temperature differences between cities and their rural surroundings, known as the urban heat island (UHI) effect (MetLink – Royal Meteorological Society Urban Heat Islands). This urban-rural contrast is a result of city diversity, including increased pervious surfaces, reflective materials, and deep canyon structures that trap heat close to the surface. These can all increase local temperatures and influence people’s thermal comfort. 

It is a challenge to represent these effects in models as city geometry is so complicated. Additionally, the low resolution of numerical weather prediction (NWP) models makes it impossible to simulate individual buildings and streets. So, we make simplifications, with one common approach assuming that streets are infinitely long and of a constant width, with equal-height buildings – an ‘infinite street canyon’. Although this could be a good assumption for suburbs, it may not be representative of the complex structure of larger cities. 

Our work focuses on urban radiation, as the amount of the sun’s energy a surface absorbs and reflects controls the other urban processes. To accurately simulate urban areas and their exchanges we need information about their structure, but there is a lack of global data on urban morphology. Additionally, we need more computationally efficient ways of describing urban energy exchanges in models. Recent model developments are moving towards multi-layer urban canopy descriptions, allowing realistic effects i.e., shadowing of short buildings by taller ones. One example for urban radiation is ‘SPARTACUS-Surface’ (GitHub – ecmwf/spartacus-surface: Radiative transfer in forests and cities) which requires profiles of building cover and wall area with height. 

The main errors that arise when modelling urban radiation are from: the radiation scheme itself; determining the city morphology from a few parameters; and knowing the exact urban parameters for each city. Previously, our work quantified the first for solar radiation (Evaluation of the SPARTACUS-Urban Radiation Model for Vertically Resolved Shortwave Radiation in Urban Areas | SpringerLink). Our new paper aimed to quantify the second (Characterising the vertical structure of buildings in cities for use in atmospheric models – ScienceDirect). 

To achieve this, we identified and parameterised urban morphology profiles, with a focus on those needed for SPARTACUS-Surface – through determination of coefficients and methods that hold for multiple countries worldwide, covering the range of urban variability both between and within cities. We studied the morphology of six cities worldwide using building height data at a 2 km × 2 km resolution: Auckland (New Zealand), Berlin (Germany), Birmingham (UK), London (UK), New York City (USA), and Sao Paulo (Brazil). The main parameters we used in the work were the cover of buildings at the surface, the mean building height, and the wall area. 

Urban morphology parameters derived at 2 km × 2 km resolution for six cities (Adapted from Stretton et al. 2023)

The parameterisations developed have different complexity levels, with decreasing input data requirements, allowing us to identify which level of data is required before a difference in the results. To parameterise the building cover with height, we use the mean building height and the surface building cover. The profiles of building wall area are parameterised using an ‘effective building diameter’. This assumes that the building cover and building wall area are proportional to each other, describing the width of buildings at each height if they were identical cubes or cylinders. We find that this can be roughly assumed to be 20 m across all cities.

The impact of the relations for city structure that we developed had on the radiation fluxes were tested using SPARTACUS-Surface, focusing on the top of canopy albedo, and the absorbed radiation. The study revealed that we can determine the vertical structure of any urban area assuming we know three simple characteristics (surface building cover, mean building height, and an effective building diameter of 20 m), with errors for albedo up to 10%. This is improved to 2% when using a better effective building diameter, calculated from the exact wall area.

This work shows that there are skillful and efficient ways to characterize cities for computationally expensive NWP models. These findings are even more useful and applicable as we move to the next-generation of models that resolve the vertical structure of cities. Also, this work reflects the need for large-scale datasets to communicate the variability of cities form and materials, which are required for these parameterisation approaches. Particularly here, we show the need for datasets of building cover and mean building height.

References:

Harman, M. J. Best, and S. E. Belcher, 2004: Radiative exchange in an urban street canyon. Boundary-Layer Meteorol., 110, 301–316, https://doi.org/10.1023/A:1026029822517.

Heaviside, C., H. Macintyre, and S. Vardoulakis, 2017: The Urban Heat Island: Implications for Health in a Changing Environment. Curr. Environ. Heal. reports, 4, https://doi.org/10.1007/s40572-017-0150-3.

Hogan, R. J., 2019a: Flexible Treatment of Radiative Transfer in Complex Urban Canopies for Use in Weather and Climate Models. Boundary-Layer Meteorol., https://doi.org/10.1007/s10546-019-00457-0.

Hogan, R. J., 2021: spartacus-surface. GitHub Repos.,.

Lindberg, F., and C. S. B. B. Grimmond, 2011b: Nature of vegetation and building morphology characteristics across a city: Influence on shadow patterns and mean radiant temperatures in London. Urban Ecosyst., 14, 617–634, https://doi.org/10.1007/s11252-011-0184-5.

Mccarthy, M. P., M. J. Best, and R. A. Betts, 2010: Climate change in cities due to global warming and urban effects. Geophys. Res. Lett., https://doi.org/10.1029/2010GL042845.

Meehl, G. A., and C. Tebaldi, 2004: More intense, more frequent, and longer lasting heat waves in the 21st century. Science, 305, 994–997, https://doi.org/10.1126/SCIENCE.1098704.

Oke, T. R., G. Mills, A. Christen, and J. A. Voogt, 2017: Urban climates.

Stretton, M. A., W. Morrison, R. J. Hogan, and S. Grimmond, 2022: Characterising the vertical structure of buildings in cities for use in atmospheric models. Urban Climate, https://doi.org/10.1016/j.uclim.2023.101560.

Stretton, M. A., R. J. Hogan, S. Grimmond, and W. Morrison, 2023: Evaluation of the SPARTACUS-Urban Radiation Model for Vertically Resolved Shortwave Radiation in Urban Areas. Boundary-Layer Meteorol., 184, 301–331, https://doi.org/10.1007/s10546-022-00706-9.

Yang, X., and Y. Li, 2015: The impact of building density and building height heterogeneity on 257 average urban albedo and street surface temperature. Build. Environ., 90, 146–156.

 

Posted in Climate modelling, Urban meteorology | Leave a comment

Rapid developing, severe droughts will become more common over the 21st Century

By: Emily Black

At the height of the 2012 corn growing season, two thirds of the United States was hit by a sudden drought. The photographs below compare 2012 to a normal year:  

Phenocam images taken at MOISST, which is adjacent to the Marena mesonet station, on (a) 1 Jul 2012, (b) 11 Aug 2012, (c) 1 Jul 2014, and (d) 11 Aug 2014. All images were taken at 10:30 local time. Otkin et al. 2018 https://journals.ametsoc.org/view/journals/bams/99/5/bams-d-17-0149.1.xml

Earlier this year, a similarly sudden drought dried out grasslands in Hawaii, contributing to the wildfires that devastated Maui.

There is a mounting body of evidence indicating that such ‘flash droughts’ are becoming more frequent and intense due to climate change, as discussed in this study. Consequently, understanding the factors driving flash droughts in current and future climates has become an increasingly urgent concern. 

Recent research conducted at the University of Reading and the National Centre for Atmospheric Science has shed light on this issue. The findings show that flash droughts are consistently preceded by anomalously low relative humidity and precipitation. Interestingly, the study suggests that heat waves do not cause flash droughts, although flash droughts can cause heat waves. 

Over the next century, flash droughts are projected to become more common globally. The plot below shows the percentage change in flash drought occurrence compared to 1960-2100, under a range of shared socioeconomic pathways: 

The most severe changes are projected in Europe, the continental US, eastern Brazil and southern Africa: 

To find out more, have a look at my paper in Advances in Atmospheric Science: http://www.iapjournals.ac.cn/aas/en/article/doi/10.1007/s00376-023-2366-5 

Posted in Climate, Climate change | Leave a comment

More severe wet and dry extremes as rapid warming of climate continues

By: Professor Richard Allan

The UK weather has recently been characterised by large swings between wet and dry periods and with record heat this June and September. Globally, this September was the warmest on record and 2023 is set to be the warmest year on record and will be remembered for the hot, wet and dry weather extremes including also severe wildfires. And as the climate continues to warm, the extremes of wet and dry will further intensify.

Flooding on a street in Whitley; Image via dachalan on Flickr (CC BY-NC-SA 2.0).

Image via dachalan on Flickr (CC BY-NC-SA 2.0 – https://creativecommons.org/licenses/by-nc-sa/2.0/).

In new research published in Environmental Research Letters, satellite and ground-based data are combined with simulations since the 1950s to show that the range between the wettest and driest time of the year is growing as the climate warms.

This work looks in detail at the difference between the amount of water arriving at the surface from precipitation such as rain and snow and the amount leaving due to evaporation. This is important in affecting how much water people and plants can use. If there is too much building up, flooding can occur but when there is a lack of rain, the soils can dry and eventually lead to drought conditions.

The new analysis finds that the global water cycle is becoming more intense. Wet times of the year, when precipitation is much more than evaporation, are becoming even wetter, but periods of drying, when evaporation can be larger than precipitation, are also becoming more intense.

For every degree Celsius of global warming, the difference between precipitation and evaporation in the wettest and driest times of the years is becoming larger by about 3 or 4 percent. This means there is a larger contrast between wet and dry spells.

In some regions such as northern North America and northern Eurasia, the contrast between wet and dry is expected to increase more than 20% by the end of this century (see diagram).

A world map diagram which shows increasing range between the wettest and driest time of the year by the end of the twenty-first century, in percent. The contrasts between wet and dry times of the year increases by more than 20% in some regions.

This is important since it means that ensuring a reliable availability of fresh water becomes an increasing challenge but also because the most damaging wet and dry seasons in a year will become more dangerous.

Patterns of future change are also found to resemble present day trends, which adds to evidence for a more variable and extreme water cycle as the climate continues to warm.

It may seem strange that we could get more extreme dry and more extreme wet spells as the climate warms, but this is possible because a warmer atmosphere is a thirstier atmosphere – it can more effectively sap the soil of its moisture in one region and dump this extra water as heavy rainfall in storms and monsoons, increasing the contrast in weather between regions and between different times of the year.

This increasing contrast can lead to severe consequences, such as more intense flooding during wet periods and more rapidly developing droughts as dry spells take hold.

Rapid swings between drought and severe flooding are known to be particularly difficult for countries to deal with. Recent research published in Advances in Atmospheric Sciences by Professor Emily Black has shown that the frequency of “flash” droughts are projected to more than double in many regions over the twenty-first century. These types of rapidly developing droughts can damage crops and will likely become more frequent in parts of the world including South America, Europe, and southern Africa.

As our greenhouse gas emissions continue to heat the planet, there will be greater swings between drought and deluge conditions that will become more severe over time.

We have already seen severe flooding in Japan, China, South Korea and India in 2023, which has caused deaths, damage and power cuts.

It is only with rapid and massive cuts in greenhouse gas emissions that we can limit warming and the increasing severity of wet and dry spells. Understanding these changes is vital for planning and managing our water resources, as well as improving predictions of how the water cycle will evolve in a warming world.

Richard Allan is Professor of Climate Science at the University of Reading.

Posted in Climate, Climate change | Leave a comment

“…since records began” – Christopher Wren’s first automatic weather station

We restart the weekly blog with a contribution from Professor Giles Harrison. With the blog being down over the summer, Giles‘ contribution was posted on Professor Maarten Ambaum’s excellent blog, where we direct readers until regular service resumes next week.

https://readingphysics19265874.wordpress.com/2023/07/28/since-records-began-christopher-wrens-first-automatic-weather-station/

The tercentenary of Wren’s death this year is being marked with a series of events, including an exhibition on his ideas about Sign Language, Beehives, Anaesthesia, Astronomy, Microscopy, Urban Design, Sheltered Living and Weather Recording, at the Old Royal Naval College at Greenwich. The exhibition continues until 12 Nov 2023, https://ornc.org/whats-on/christopher-wren-what-legacy-now/

Posted in Climate | Leave a comment

How to improve a climate model: a 24-year journey from observing melt ponds to their inclusion in climate simulations

By: David Schroeder

Melt ponds are puddles of water that form on top of sea ice when the snow and ice melts (see Figure). Not all the water drains immediately into the ocean, but it can stay and accumulate on top of the sea ice for several weeks or months (Ref: https://blogs.reading.ac.uk/weather-and-climate-at-reading/2017/melt-ponds-over-arctic-sea-ice/

Figure: Melt ponds on sea ice (Credit: Don Perovich)

A momentous field campaign was carried out in 1998 on the Arctic sea ice: the Surface Heat Budget of the Arctic Ocean (SHEBA) experiment (https://www.nsf.gov/pubs/2003/nsf03048/nsf03048_3.pdf) – a role model for the latest and largest Arctic expedition MOSAIC in 2019/2020 (https://mosaic-expedition.org/expedition/). One aim was to understand and quantify the sea ice-albedo feedback mechanism on scales ranging from meters to thousands of kilometers. The differences in albedo (fraction of shortwave radiation reflected at the surface and, thus, not used to heat the surface) between snow-covered sea ice (~85%), bare sea ice (~60-70%), ponded sea ice (~30%) and open water (<10%) are huge and cause the most important feedback for sea ice melt: The more and the earlier snow and ice melts, the larger the pond and open water fraction, the more shortwave radiation will be absorbed increasing the melting. Melt ponds play an important part in the observed reduction and thinning of Arctic sea ice during last decades.

Continuous SHEBA measurements over the whole melt season in 1998 allowed the development of models representing the melting cycle: from the onset of melt pond formation, spreading, evolution and drainage over late spring and summer, towards freeze-up in the late summer and autumn. Starting with a one-dimensional heat balance model (Taylor and Feltham, 2004), it took about 10 years to develop a pond model suitable for a Global Climate Model (GCM) (Flocco et al., 2010; 2012). Melt pond formation is controlled by small-scale sea ice topography. This is not available in a GCM with coarser resolution. However, we could use the sub-gridscale ice thickness distribution (5 different ice thickness categories for each grid cell) as a proxy for topography and simulate the evolution of pond fraction assuming melt water runs from the thicker ice to the thinner ice. With further adjustments to the albedo scheme (Ridley et al., 2018), the pond model could finally be used in the UK Climate Model HadGEM3. The HadGEM3 simulations for the latest IPPC report (https://www.ipcc.ch/report/ar6/wg2/) include our pond model.

What is the impact of the melt pond model on the performance of the HadGEM3 simulations? It is noteworthy that HadGEM3  has a stronger climate sensitivity (global warming with respect to CO2 increase) compared to its predecessor HadGEM2  or most other climate models (Mehl et al., 2020). But is this due to the melt ponds? Lots of model components were changed at the same time, so it is impossible to specify the individual impact. To address this, Diamond et al. (2023) carried out HadGEM3 simulations with 3 configurations which only differ with respect to melt pond treatment (our pond scheme, simple albedo tuning to account for the impact of melt ponds and no melt ponds). Historical or future projections would require an ensemble simulation to distinguish between internal variability and impact of pond scheme. Thus, 100 year long constant forcing simulations have been chosen.

While Arctic sea ice results between the simple albedo tuning and our full pond scheme do not differ significantly for pre-industrial conditions, the impact on near future conditions are remarkable: The simple tuning never yields an ice-free summer Arctic, whilst our pond scheme yields an ice-free Arctic 35% of years and raises autumn Arctic air temperatures by 5 to 8 °C.  Thus, the pond treatment has a large impact on projections when the Arctic will become ice-free. This is a striking example of the impact

References:

Diamond, R., Schroeder, D., Sime, L.C., Ridley, J., and Feltham, D.L.: Do melt ponds matter? The importance of sea-ice parametrisation during three different climate periods. J. of Climate, under review.

Flocco, D., D. L. Feltham, and A. K. Turner, 2010: Incorporation of a physically based melt pond scheme into the sea ice component of a climate model. Journal of Geophysical Research: Oceans, 115 (C8).

Flocco, D., D. Schroeder, D. L. Feltham, and E. C. Hunke, 2012: Impact of melt ponds on arctic sea ice simulations from 1990 to 2007. Journal of Geophysical Research: Oceans, 117 (C9).

Mehl, G. A., C. A. Senior, V. Eyring, G. Flato, J.-F. Lamarque, R. J. Stouffer, K. E. Taylor, and M. Schlund, 2020: Context for interpreting equilibrium climate sensitivity and transient climate response from the cmip6 earth system models. Science Advances, 6 (26).

Ridley, J. K., E. W. Blockley, A. B. Keen, J. G. Rae, A. E. West, and D. Schroeder, 2018b: The sea ice model component of hadgem3-gc3. 1. Geoscientific Model Development, 11 (2), 713–723.

Taylor, P., and D. Feltham, 2004: A model of melt pond evolution on sea ice. Journal of Geophysical Research: Oceans, 109 (C12).

Posted in Arctic, Climate modelling, Cryosphere, IPCC, Numerical modelling, Polar | Leave a comment

Cycling In All Weathers

By: David Brayshaw

In a few weeks’ time, I’ll be taking some time off for an adventure: spending 3-weeks cycling the entire 3,400 km of this year’s Tour de France (TdF) route.  I’ll be with a team riding just a few days ahead of the professional race, aiming to raise £1M for charity.  Although this is a purely personal challenge – unrelated to my day job here in the department – being asked to write this blog set me thinking about the connections between cycling and my own research in weather and climate science.

Weather is obviously important to anyone cycling outdoors: be it extremes of rain, wind or temperature.  Cycling in the rain can be miserable but, more than that, it can lead to accidents on slippery roads and poor visibility for riders.   Cold temperatures and wind chill pose challenges particularly when descending at speeds of up to 50 mph in the high mountains (in years gone by professional cyclists often took a newspaper from a friendly spectator at the top of a climb to shove it down the front of their cycling jersey to protect themselves from the worst of the wind chill).  Air resistance and wind play a major role more generally: the bunching up of the peloton occurs as riders save energy by staying out of the wind and riding close behind the cyclist in front.  While, while headwinds sap riders’ energy and lower their speed, it’s crosswinds that blow races apart.  In that situation, the wind-shielding effect runs diagonally across the road, shredding the peloton into diagonal lines as riders fight for position and cover.

Photo: Grim conditions on a training ride in the Yorkshire Wolds, April 2023.

Last year’s TdF race, however, took place in a heat wave.  The athletes did their work in air temperatures approaching 40 oC, stretching the limits of human performance in extreme temperatures.  On some days the roads were sprayed with water to stop the tarmac melting (road temperatures were often closer to 60 oC), and extreme weather protocols were called upon (potential adjustments include changes to the start time or route, making more food and water available, even cancelling whole stages).  All this comes with risks and costs (human, environmental, financial) for a range of people and organisations (the riders and spectators; the organisers and sponsors; and the towns and communities the ride goes through).  Moreover, heatwaves can only be expected to become more common in the years to come.

From a meteorological perspective, the “good news” is that tools are available to help quantify, understand and manage weather risks.  High-quality short-range (hours to days) forecasting is obviously essential during the event itself but subseasonal to seasonal (S2S) forecasts or longer-term climate change projections may also help to manage risk over a longer horizon (e.g., hire of water trucks, anticipating the need for route modification, use of financial products to mitigate losses if stages are cancelled or adjusted, even reconsidering the timing of the event itself if July temperatures become intolerable in the decades to come).

The specifics of the decisions and consequences described here for this particular race are simply speculation on my part (I have not done any in-depth research on climate services for cycling!).  However, the nature of the “climate impact problem” should be familiar to anyone working in the field.  As an example, some recent work I was involved in which produced a proof-of-concept demonstration of how weeks-ahead forecasts could be used to improve fault management and maintenance scheduling in telecommunications (see figure below and full discussion here), but many more examples can be found (see here for a recent review).  In such work, there are usually two core challenges.  Firstly, to link quantitative climate data (say, skillful probabilistic predictions of air temperature weeks ahead) with the impact of concern (say, the need to cancel part of a stage and the financial losses incurred by the host town that is then not visited).  Then, secondly, to identify the mitigating actions that can take place (say, the purchase of insurance or a financial hedge) and a strategy for their uptake (say, a decision criteria for when to act and at what cost).  The broad process is discussed in two online courses offered here in the department (“Climate Services and Climate Impact Modelling” and “Climate Intelligence: Using Climate Data to Improve Business Decision-Making”).

Figure: Use of week-ahead sub-seasonal forecasts to anticipate and manage line faults.  Left panel demonstrates that predictions of weekly fault rates made using a version of ECMWF’s subseasonal forecast system (solid and dashed lines represent two different forecast methods) outperform predictions made using purely historic “climatological” knowledge (dotted line).  The right panel illustrates the improved outcomes possible with the improving forecast information (from red to purple to blue curves): i.e., by using a “better” forecast it is possible to achieve either higher performance for the same resources, or the same performance for fewer resources (here as an illustrative schematic but an application to “real” data is available in the cited paper).  Figures adapted from or based upon Brayshaw et al (2020, Meteorological Applications), please refer to the open-access journal article for detailed discussion.

For this summer, however, I’m just hoping for good weather for my ride.  Thankfully I won’t be trying to “race” the distance (merely survive it!), so a mix of not too hot, not too wet, not too windy would just be perfect.  With a bit of luck, hopefully, I’ll make it all the way from the start line in Bilbao to the finish in Paris!

If you’d like to find out more about my ride or the cause I’m supporting then please visit my personal JustGiving page (https://www.justgiving.com/fundraising/david-brayshaw-tour21-2023).

References:

  • Brayshaw, D. J., Halford, A., Smith, S. and Kjeld, J. (2020) Quantifying the potential for improved management of weather risk using subseasonal forecasting: the case of UK telecommunications infrastructure.Meteorological Applications, 27 (1). e1849. ISSN 1469-8080 doi: https://doi.org/10.1002/met.1849

  • White, C. J., Domeisen, D. I.V., Acharya, N., Adefisan, E. A., Anderson, M. L., Aura, S., Balogun, A. A., Bertram, D., Bluhm, S., Brayshaw, D. J. , Browell, J., Büeler, D., Charlton-Perez, A., Chourio, X., Christel, I., Coelho, C. A. S., DeFlorio, M. J., Monache, L. D., García-Solórzano, A. M., Giuseppe, F. D., Goddard, L., Gibson, P. B., González, C. R., Graham, R. J., Graham, R. M., Grams, C. M., Halford, A., Huang, W. T. K., Jensen, K., Kilavi, M., Lawal, K. A., Lee, R. W., MacLeod, D., Manrique-Suñén, A., Martins, E. S. P. R., Maxwell, C. J., Merryfield, W. J., Muñoz, Á. G., Olaniyan, E., Otieno, G., Oyedepo, J. A., Palma, L., Pechlivanidis, I. G., Pons, D., Ralph, F. M., Reis, D. S., Remenyi, T. A., Risbey, J. S., Robertson, D. J. C., Robertson, A. W., Smith, S. , Soret, A., Sun, T. , Todd, M. C., Tozer, C. R., Vasconcelos, F. C., Vigo, I., Waliser, D. E., Wetterhall, F. and Wilson, R. G. (2022) Advances in the application and utility of subseasonal-to-seasonal predictions. Bulletin of the American Meteorological Society, 103 (6). pp. 1448-1472. ISSN 1520-0477 doi: https://doi.org/10.1175/BAMS-D-20-0224.1

Posted in Climate Services, Environmental hazards, Seasonal forecasting, subseasonal forecasting | Leave a comment

Flying Through Storms To Understand Their Interaction with Sea Ice: The Arctic Summer-time Cyclones Project and Field Campaign

By: Ambrogio Volonté

Arctic cyclones are the leading type of severe weather system affecting the Arctic Ocean and surrounding land in the summer. They can have serious impacts on sea-ice movement, sometimes resulting in ‘Very Rapid Ice Loss Events’, which present a substantial challenge to forecasts of the Arctic environment from days out to a season ahead. Summer sea ice is becoming thinner and more fractured across widespread regions of the Arctic Ocean, due to global warming. As a result, winds can move the ice around more easily. In turn, the uneven surface can exert substantial friction on the atmosphere right above it, impacting the development of weather systems. Thus, a detailed understanding of the two-way relationship between sea ice and Arctic cyclones is crucial to allow weather centres to provide reliable forecasts for the area, an increasingly important issue as the Arctic sees growing human activity.

This is the main goal of the Arctic Summer-time Cyclones project, led by Prof John Methven and funded by the UK Natural Environment Research Council (NERC). To this end, we designed a field campaign aiming to fly into Arctic cyclones developing over the marginal ice zone (that is the transitional area between pack ice and open ocean, where the ice is thinner and fractured, and where leads and melt ponds can be present). The campaign was based in Svalbard (Norwegian Arctic) and took place in July and August 2022, one year later than originally planned due to the Covid pandemic. The field campaign team included scientists from the University of Reading (John Methven, Suzanne Gray, Ben Harvey, Oscar Martinèz-Alvarado, Ambrogio Volonté and Hannah Croad), the University of East Anglia (UEA), and the British Antarctic Survey (BAS). We were joined by researchers from the US and France, funded by the Office of Naval Research (USA).

Figure 1: Some components of the Arctic Summer-time Cyclones field campaign team in front of the Twin Otter aircraft. Photo by Dan Beeden (BAS).

Using the BAS MASIN Twin Otter aircraft, we performed 15 research flights during the campaign, targeting four Arctic cyclones and several other weather features associated with high winds near the surface. Flying at very low levels (even below 100ft when allowed by visibility conditions and safety standards) we were able to detect the turbulent fluxes of heat and momentum characterising the interaction between surface and atmosphere. Vertical profiles and stacks of horizontal legs at different heights were used to sample for the first time the 3D structure of wind jets present in the first km above the surface in Arctic summer cyclones. Our partners from France and US also completed a similar number of flights using their SAFIRE ATR42 aircraft. Although their activities were mainly focused on cloud structure and mixed phase (ice-water) processes higher up, some coordinated flights were carried out, with both aircrafts flying in the same area to maximise data collection. For more details on our campaign activities (plus photos and videos from the Twin Otter!) see the ArcticCyclones Twitter account and the blogs on our project website.

Figure 2: An example of sea ice as seen from the cockpit of the Twin Otter during the flight on 30 July 2022. Photo by Ian Renfrew (UEA).

Now that the field campaign has concluded, data analysis is proceeding apace. Flight observations are being compared against model data from operational weather forecasts and dedicated high-resolution simulations. While our colleagues at the University of East Anglia are analysing the observed turbulent fluxes over sea ice to improve their representation in forecast models, here at Reading we are looking at the detailed 3D structure of Arctic cyclones and at the processes driving their lifecycle. Preliminary results highlight the sharpness of the low-level wind jet present in their cold sector, with observations suggesting that jet cores are stronger and shallower than shown by current models. However, more detailed analysis is still needed to confirm these results. At the same time, novel analysis methods are being implemented on experimental model data, taking advantage of the properties of conservation and inversion of atmospheric variables such as potential vorticity and potential temperature. The aim is to isolate the contributions of individual processes, such as friction and heating, to the dynamics of the cyclone and thus highlight the effects of atmospheric-surface interaction on cyclone development.

Figure 3: Example of flight planner map (software developed by Ben Harvey, Reading) used to set up the flight route of one of the campaign flights. Background data from UK Met Office (Crown copyright).

While we are surely missing the sense of adventure of our Arctic field campaign, the excitement for the scientific challenge is still accompanying us as we analyse the data here in Reading and collaborate with our UK and international partners. Stay tuned if you are interested in how Arctic cyclones work, how they interact with the changing sea ice and how Arctic weather forecast can be improved. Results might soon be coming your way!

 

Posted in Arctic, Climate, Climate change, Data collection, extratropical cyclones | Leave a comment

Two Flavours of Ocean Temperature Change and the Implication for Reconstructing the History of Ocean Warming

Introducing Excess and Redistributed Temperatures. 

By: Quran Wu

Monitoring and understanding ocean heat content change is an essential task of climate science because the ocean stores over 90% of extra heat that is trapped in the Earth system. Ocean warming results in sea-level rise which is one of the most severe consequences of anthropogenic climate change.

Ocean warming under greenhouse gas forcing is often thought of as extra heat being added to the ocean surface by greenhouse warming and then carried to depths by ocean circulation. This one-way heat transport diagram assumes that all subsurface temperature changes are due to the propagation of surface temperature changes, and is widely used to construct conceptual models of ocean heat uptake (for example, the two-layer model in Gregory 2000).

Recent studies, however, have found that ocean temperature change under greenhouse warming is also affected by a redistribution of the original temperature field (Gregory et al. 2016). The ocean temperature change due to the redistribution is referred to as redistributed temperature change, while that due to propagation of surface warming is referred to as excess temperature change.

A Dye Analogy

To help explain the separation of excess and redistributed temperature, let us consider a dye analogy. Heating the ocean from the surface is like adding a drop of dye into a glass of water that already has a non-uniform distribution of the same dye. After the dye injection, two things happen simultaneously. First, the newly-added dye gradually spreads into the water in the glass (excess temperature). Second, the dye injection disturbs the water and causes water motion that rearranges the original dye (redistributed temperature). Both processes contribute to changes in dye concentrations.

Climate Model Simulation

Figure 1: Time evolution of global-mean ocean temperature change (in Kelvin) under increasing greenhouse gas emission in a climate model simulation (a). Change in (a) is decomposed into excess temperature change (b) and redistributed temperature change (c).

Excess and redistributed temperatures are both derived from thought experiments; neither of them can be directly observed in the real world. Here, we demonstrate their behaviours using a climate model simulation under increasing greenhouse gas emission. The simulation shows that ocean warming starts from the surface, and propagates into depths gradually, reaching 500 m after 50 years (Figure 1a). The ocean warming is mostly driven by excess temperature change (compare Figures 1a with 1b) but strongly disrupted by a downward heat redistribution near the surface (cooling at the surface and warming underneath) (Figure 1c). The downward heat redistribution is caused by a reduction of ocean convection (which pumps heat upward), because surface warming stabilises water columns.

Implications

Distinguishing excess from redistributed temperature change is important because they behave in different ways. While one can reconstruct excess temperature at depths by propagating its surface change using ocean transports, the same cannot be done with redistributed temperature. This is because temperature redistribution can potentially happen anywhere in the ocean, unlike extra heat, which can only enter the ocean from the surface (under greenhouse warming). Such a distinction has important implications for estimating the history of ocean warming from surface observations.

Ocean warming is traditionally estimated by interpolating in-situ temperature measurements, which were gathered in discrete locations and times, to the global ocean. This in-situ method suffers a large uncertainty because the ocean remains poorly sampled until the global deployment of Argo floats (a fleet of robotic instruments) in 2005.

A new approach to estimate ocean warming is to propagate its surface signature, that is sea surface temperature change, downward using information of ocean transports (Zanna et al. 2019). This transport method is useful because it relies on surface observations, which have a longer historical coverage than subsurface observations. However, this method ignores the fact that part of surface temperature change is due to temperature redistribution, which does not correspond to subsurface temperature change. In a computer simulation of the historical ocean, we found that propagating sea surface temperature change results in an underestimate of simulated ocean warming due to redistributive cooling at the surface (as shown in Figure 1c) (Wu and Gregory 2022). This result highlights the need for isolating excess temperature change from surface observations when applying the transport method to reconstruct ocean warming.

Acknowledgements

Thanks to Jonathan Gregory for reading an early version of this article and providing useful comments and suggestions.

References:

Gregory, J. M., 2000: Vertical heat transports in the ocean and their effect on time-dependent climate change. Climate Dynamics, 16, 501–515, https://doi.org/10.1007/s003820000059.

Gregory, J. M., and Coauthors, 2016: The Flux-Anomaly-Forced Model Intercomparison Project (FAFMIP) contribution to CMIP6: investigation of sea-level and ocean climate change in response to CO2 forcing. Geoscientific Model Development, 9, 3993–4017, https://doi.org/10.5194/gmd-9-3993-2016.

Wu, Q., and J. M. Gregory, 2022: Estimating ocean heat uptake using boundary Green’s functions: A perfect‐model test of the method. Journal of Advances in Modeling Earth Systems, 14, https://doi.org/10.1029/2022MS002999.

Zanna, L., S. Khatiwala, J. M. Gregory, J. Ison, and P. Heimbach, 2019: Global reconstruction of historical ocean heat storage and transport. Proceedings of the National Academy of Sciences, 116, 1126–1131, https://doi.org/10.1073/pnas.1808838115.

 

Posted in Climate, Climate change, Climate modelling, Oceans | Leave a comment

Using Old Ships To Do New Science

By: Praveen Teleti

Weather Rescue at Sea: its goals and progress update.

Observing the environment around us is fundamental to learning about and understanding the natural world. Before the Renaissance, everyday weather was thought to be works of divine or supernatural hence beyond human comprehension. Trying to understand the weather was considered so futile that an indecisive or fickle-minded person was called weather-cock, who could turn any way without any reason. In some quarters, efforts to hypothesise rules of atmosphere, let alone forecast the weather, was considered heretical and blasphemous. 

 However, weather played a significant role in day-to-day life from timings of sowing and harvesting, well-being of cattle and other domesticated animals, trade-commerce, even outcomes of conflicts. The treatise written on weather by Greek philosopher, Aristotle in 340 BC was forgotten, and no gains were made on the understanding of the subject until 17th-18th Century. The weather phenomena was too abstract to comprehend without systematic accumulation of weather observations, and it became possible only after invention of weather instruments. Figure 1: The average number of observations recorded per month for each year in the ICOADS (International Comprehensive Ocean-Atmosphere Data Set) dataset, the sizes of data points are proportional to the percent of oceans covered by observations that year. 

Due to the precarious nature of life on sea, mariners started observing and recording weather several times a day, as recognising potential tempests in the vicinity and moving away could save their ship and their lives. Taking precautionary actions also made commercial sense in reducing loss or damage to the goods during transit. Ship owners and insurance providers encouraged and later mandated that weather observations be taken and recorded in an orderly fashion as to derive long-term benefit out of it.  

Sharing weather information was beneficial to all ships irrespective of nationalities, or the nature of companies operating them. However, by then no one common method or units of measuring the weather existed, which made the observations from different ships incompatible. To solve such a problem of incompatibility of information, a maritime conference in Brussels took place between major European powers in 1854.  

In the maritime conference of 1854, it was proposed to standardise methods of observation taking and keeping of logbooks, this led to an increase in the number of usable observations from 1854 onwards. About the same time, the sinking of the Royal Charter ship in a storm off the north coast of Anglesey in October 1859 inspired Vice-Admiral Robert FitzRoy to develop weather charts which he described as “forecasts”, thus the Met Office was born. He used the telegraphic network of weather stations around the British Isles to synthesise the current state of weather.  

There is a scientific interest in understanding the climate of the early industrial era against which our present climate could be measured. Invaluable data from many hundreds of thousands of such ship journeys can be used to inform and to estimate the changes that occurred over many decades. Data rescue (transcribing hand-written observations into computer readable digital format) of historical logbooks has been taking place for decades, but to manually transcribe an almost inexhaustible number of logbooks by individual researchers, would take thousands of human lifetimes. 

As a result, large gaps have remained in our knowledge of the climate, both in space and time. The 19th Century has fewer observations available than the 20th Century in the world’s largest observation meteorological dataset, ICOADS version 3 (International Comprehensive Ocean-Atmosphere Data Set, Freeman et al. 2017). On closer inspection, the average number of monthly observations and percent of global coverage in the 1860s and 1870s is poor compared to other decades after 1850 (Figure 1). 

With this context, the Weather Rescue At Sea project was launched to use the citizen science-based Zooniverse platform to recover some of these observations and make them usable, with a focus on ships travelling through the Atlantic, Indian and Pacific Ocean basins in the 1860s and 1870s. Filling in the gaps in our knowledge will remove ambiguity in how the climate varied historically in many regions where observations are currently poor or non-existent. 

The data generated through this project will help fill many crucial gaps in the large climate datasets (e.g., ICOADS) which will be used to generate new estimates of the industrial and pre-industrial era baseline climate. But more generally, this data and data from other historical sources are used to improve the models and reanalysis systems used for climate and weather research. We need your help to data-rescue these weather observations so that scientists can analyse these observations to better understand changes in the climate since and forecast changes in the future. 

Figure 2: Ship tracks of some of the ships recovered through WRS data-rescue project 

Progress so far: Out of 248 ship logbooks used for this project, 213 logbooks are more than 80% finished, while 35 logbooks are complete. Meaning all positional and meteorological observations (e.g., Sea-level pressure, Air Temperature, Sea water Temperature, Wind speed-direction) in 35 logbooks have been transcribed (Figure 2). To date more than two million dates, positions and weather observations have been transcribed. 

We need your help to get this project across the finish line, let us give a final push to complete all logbooks. Check the poster below to volunteer. 

References:

Freeman, E., S.D. Woodruff, S.J. Worley, S.J. Lubker, E.C. Kent, W.E. Angel, D.I . Berry, P. Brohan, R. Eastman, L. Gates, W. Gloeden, Z. Ji, J. Lawrimore, N.A. Rayner, G. Rosenhagen, and S.R. Smith, 2017: ICOADS Release 3.0: A major update to the historical marine climate record. Int. J. Climatol. (CLIMAR-IV Special Issue), 37, 2211-2237 (doi:10.1002/joc.4775).

Posted in Climate, Data collection, Data rescue, Historical climatology, Reanalyses | Leave a comment

Including Human Behaviour in Models to Understand the Impact of Climate Change on People

By Megan McGrory

In 2020 56% of the global population lived in cities and towns, and they accounted for two-thirds of global energy consumption and over 70% of CO2 emissions. The share of the global population living in urban areas is expected to rise to almost 70% in 2050 (World Energy Outlook 2021). This rapid urbanization is happening at the same time that climate change is becoming an increasingly pressing issue. Urbanization and climate change both directly impact each other and strengthen the already-large impact of climate change on our lives. Urbanization dramatically changes the landscape, with increased volume of buildings and paved/sealed surfaces, and therefore the surface energy balance of a region. The introduction of more buildings, roads, vehicles, and a large population density all have dramatic effects on the urban climate, therefore to fully understand how these impacts intertwine with those of climate change, it is key to model the urban climate correctly.

Modelling an urban climate has a number of unique challenges and considerations. Anthropogenic heat flux (QF) is an aspect of the surface energy balance which is unique to urban areas. Modelling this aspect of urban climate requires input data on heat released from activities linked to three aspects of QF: building (QF,B), transport (QF,T) and human/animal metabolism (QF,M). All of these are impacted by human behaviour which is a challenge to predict, as it changes based on many variables, and typical behaviour can change based on unexpected events, such as transport strikes, or extreme weather conditions, which are both becoming increasingly relevant worries in the UK.

DAVE  (Dynamic Anthropogenic actiVities and feedback to Emissions) is an agent-based model (ABM) which is being developed as part of the ERC urbisphere and NERC APEx projects to model QF and impacts of other emissions (e.g. air quality), in various cities across the world (London, Berlin, Paris, Nairobi, Beijing, and more). Here, we treat city spatial units (500 m x 500 m, Figure 1) as the agents in this agent-based model. Each spatial unit holds properties related to the buildings and citizen presence (at different times) in the grid. QF can be calculated for each spatial unit by combining the energy emissions from QF,B, QF,T, and QF,M within a grid. As human behaviour modifies these fluxes, the calculation needs to capture the spatial and temporal variability of people’s activities changing in response to their ‘normal’ and other events.

To run DAVE for London (as a first test case, with other cities to follow), extensive data mining has been carried out to model typical human activities and their variable behaviour as accurately as possible. The variation in building morphology (or form) and function, the many different transport systems, meteorology, and data on typical human activities, are all needed to allow human behaviour to drive the calculation of QF, incorporating dynamic responses to environmental conditions.

DAVE is a second generation ABM, like its predecessor it uses time use surveys to generate statistical probabilities which govern the behaviour of modelled citizens (Capel-Timms et al. 2020). The time use survey diarists document their daily activities every 10 minutes. Travel and building energy models are incorporated to calculate QF,B and QF,T. The building energy model, STEBBS (Simplified Thermal Energy Balance for Building Scheme) (Capel-Timms et al. 2020), takes into account the thermal characteristics and morphology of building stock in each 500 m x 500 m spatial unit area in London. The energy demand linked to different activities carried out by people (informed by time use surveys) impacts the energy use and from this anthropogenic heat flux from building energy use fluxes (Liu et al. 2022).

The transport model uses information about access to public transport (e.g. Fig. 1). As expected grids closer to stations have higher percentage of people using that travel mode. Other data used includes road densities, travel costs, and information on vehicle ownership and travel preferences to assign transport options to the modelled citizens when they travel.

Figure 1: Location of tube, train and bus stations/stops (dots) in London (500 m x 500 m grid resolution) with the relative percentage of people living in that grid who use that mode of transport (colour, lighter indicates higher percentage). Original data Sources: (ONS, 2014), (TfL, 2022)

An extensive amount of analysis and pre-processing of data are needed to run the model but this provides a rich resource for multiple MSc and Undergraduate student projects (past and  current) to analyse different aspects of the building and transport data. For example, a current project is modelling people’s exposure to pollution, informed by data such as shown in Fig. 2, linked with moving to and between different modes of transport between home and work/school. Therefore the areas that should be used/avoided to reduce risk of health problems by exposure to air pollution.

Figure 2:  London (500 m x 500 m resolution) annual mean NO2 emissions (colour) with Congestion Charge Zone (CCZ, blue) and Ultra Low Emission Zone (ULEZ, pink).  Data source: London Datastore, 2022

Future development and use of the model DAVE will allow for the consideration of many more unique aspects of urban environments and their impacts on the climate and people.

Acknowledgements: Thank you to Matthew Paskin and Denise Hertwig for providing the Figures included.

References:

Capel-Timms, I., S. T. Smith, T. Sun, and S. Grimmond, 2020: Dynamic Anthropogenic activitieS impacting Heat emissions (DASH v1.0): Development and evaluation. Geoscientific Model Development, 13, 4891–4924

London Datastore, 2022: Greater London Authority, London Atmospheric Emissions Inventory 2019.

International Energy Agency, 2021: World Health Organisation 2021, (Accessed January 2023)

Liu, Y., Z. Luo, and S. Grimmond, 2022: Revising the definition of anthropogenic heat flux from buildings: role of human activities and building storage heat flux. Atmospheric Chemistry and Physics, 22, 4721–4735

ONS, 2014: Office for National Statistics, WU03UK – Location of usual residence and place of work by method of travel to work (Accessed August, 2022).

TfL, 2022: Transport for London timetables, (Accessed July 2022)

Posted in Climate, Climate change, Climate modelling, Urban meteorology | Leave a comment