Sarah Ineson's research while affiliated with Met Office and other places

What is this page?


This page lists the scientific contributions of an author, who either does not have a ResearchGate profile, or has not yet added these contributions to their profile.

It was automatically created by ResearchGate to create a record of this author's body of work. We create such pages to advance our goal of creating and maintaining the most comprehensive scientific repository possible. In doing so, we process publicly available (personal) data relating to the author as a member of the scientific community.

If you're a ResearchGate member, you can follow this page to keep up with this author's work.

If you are this author, and you don't want us to display this page anymore, please let us know.

Publications (57)


Global mean annual temperature variability. Observed temperature timeseries (1981–2023) is shown from the WMO defined definition using the average of the six leading temperature datasets. The final 2023 point is +1.45°C. The red line shows the Met Office DePreSys3 ensemble mean predictions initialized from 1st November of the preceding year and the grey (red) dots show each of the individual 40 ensemble members (for the 2024 prediction). Observed and predicted trends are calculated over 1981–2023 and quoted in the legend. The linearly detrended timeseries are also shown varying around the y = 0 axis, with the residual observed variability as pink/blue bars and the predicted variability as a red line with the 95% confidence interval. Blue and red triangles at the bottom of the plot highlight preceding winters (DJF) with an ENSO ONI index greater than ±1 K.
Key drivers of the 2024 forecast. (a) DePreSys3 forecast temperature map for 2024 relative to a 1981–2010 climatology. (b, c) Temperature anomaly map obtained from the trend and the lagged impact of a major El Niño event (the summation of the bottom two rows). (d, e) Trend assessed at each gridpoint over 1981–2023 and projected to 2024. (f, g) Estimate of the impact of a major El Niño event relative to the trend using 1997/1998 and 2015/2016 events. Numbers in brackets give the area‐weighted global mean anomaly.
Forecast temperature maps. (a) Ensemble mean DePreSys3 prediction for 2024 annual temperature relative to recent real‐time forecasts over 2016–2023. (b) The difference between the mean of the 14 members that exceed 1.5°C and the mean of the 26 members that do not. Stippling shows regions that are significantly different at the 95% level. (c, d) The coldest (c) and warmest (d) 2024 members are shown.
Hemispheric temperatures. (a, b) Northern (a) and Southern (b) hemisphere average temperature timeseries. The 2023 and 2024 years are marked by ‘+’ symbols. Observed and predicted hemispheric temperature trends are calculated over 1981–2023 and quoted in the legend. (c) The observed HadCRUT5 anomaly for 2023 (relative to 1981–2010) minus the DePreSys3 ensemble mean 2023 forecast. (d) The DePreSys3 ensemble mean 2024 forecast minus the 2023 forecast. The equator is shown as a black line in panels (c) and (d).
Will 2024 be the first year that global temperature exceeds 1.5°C?
  • Article
  • Full-text available

June 2024

·

73 Reads

Atmospheric Science Letters

Atmospheric Science Letters

·

Doug M. Smith

·

Chris Atkinson

·

[...]

·

Adam A. Scaife

Global mean near surface temperature change is the key metric by which our warming climate is monitored and for which international climate policy is set. At the end of each year the Met Office issues a global mean temperature forecast for the coming year. Following on from the new record in 2023, we predict that 2024 will likely (76% chance) be a new record year with a 1‐in‐3 chance of exceeding 1.5°C above pre‐industrial. Whilst a one‐year temporary exceedance of 1.5°C would not constitute a breach of the Paris Agreement target, our forecast highlights how close we are now to this. Our 2024 forecast is primarily driven by the strong warming trend of +0.2°C/decade (1981–2023) and secondly by the lagged warming effect of a strong tropical Pacific El Niño event. We highlight that 2023 itself was significantly warmer than the Met Office DePreSys3 forecast, with much of this additional observed warming coming from the southern hemisphere, the cause of which requires further understanding.

Download
Share

Observed and forecast temperature and precipitation for Winter (DJF) 2021/2022. (a) Ensemble mean forecast (GloSea6) screen level temperature anomaly (T2m), (b) Renalysis (ERA5; Hersbach et al., 2020) T2m anomaly as a representation of the observed anomaly, (c) Ensemble mean forecast precipitation anomaly, (d) Observed precipitation anomaly from the Global Prediction Climatology Project (GPCP; Adler et al., 2003). Hatching indicates where the magnitude is significantly greater than zero for the model ensemble mean (a) and (c), and where the observed anomaly is greater than 1 standard deviation of variability in the hindcast period (b) and (d).
Simulation of potential drivers of North Atlantic‐European winter conditions in 2021/2022. Monthly mean ensemble timeseries of (a) Niño 3.4 sea surface temperature (SST) anomaly (K) and (b) the equivalent for monthly‐mean SPV strength (m/s) compared to ERA5 data.
Predicted and observed atmospheric circulation compared to the patterns associated with remote drivers. The ensemble mean December–February MSLP anomaly in the forecast (a) is shown with shading indicating where it is significantly different from zero. Observed anomalies (b) use hatching to show where the magnitude of anomalies is greater than 1 standard deviation of variability in the hindcast period. The composite MSLP anomalies for hindcast La Niña events (c) and observed La Niña events (d) are shown for comparison. Similarly, composite anomalies for strong statospheric polar vortex winters are shown for hindcast members (e) and observations (f). La Niña events are identified as a Niño 3.4 index of more than 0.5 K below the DJF climatology. Patterns in (c–f) are scaled by the amplitude of the relevant index of each driver in the forecast.
Sub‐seasonal evolution of forecast and observed circulation patterns through winter 2021/2022. GloSea6 ensemble mean MSLP anomaly forecasts are shown for (a) December, (b) January, and (c) February. Observed anomalies from the ERA5 reanalysis (1993–2016) climatology for (d) December, (e) January and (f) February. Hatching shows areas where the ensemble mean anomaly is greater than 2 standard errors (95% confidence level) for the model and greater than one standard deviation of variability in the hindcast period for the observations.
North Atlantic‐European wind speed forecasts and storminess in winter 2021/2022. Seasonal average 10 m wind speed anomaly is shown for the ensemble mean prediction prior to winter along with the ERA5 analysis – (a) and average daily mean wind speeds from the ERA5 reanalysis during Winter 2021/2022 are shown in (b). The horizontal line represents climatological winter mean wind speed and vertical lines show the timing of named storms (see text), annotated with storm name.
Predictability of European winter 2021/2022: Influence of La Niña and stratospheric polar vortex

June 2024

·

53 Reads

Atmospheric Science Letters

Atmospheric Science Letters

The Northern Hemisphere winter of 2021/2022 exhibited a positive North Atlantic Oscillation (NAO) which led to largely mild and wet conditions for Northern Europe. A moderate La Niña in the tropical Pacific and a stronger than average stratospheric polar vortex together explained the observed anomalies over the winter. Winter 2021/2022 was well predicted in general by seasonal forecast systems. The ensemble mean indicated a positive winter NAO and the forecast spread of forecasts from the Met Office GloSea6 seasonal prediction system spanned the observed mean sea level pressure anomaly for the whole winter and the individual months. However, December showed the largest departure from the mean of the forecast which is consistent with evidence from previous work that early winter ENSO teleconnections are too weak in model predictions. Nevertheless, around one in four members captured the negative NAO pattern in December. The strong pressure gradient and positive NAO predicted for the latter part of the winter allowed successful warning of the possibility of above average storminess and strong winds which occurred in February 2022. This is potentially useful information for the energy sector who increasingly rely on wind power and the insurance industry for warning of storm damage.


Statistics of sudden stratospheric warmings using a large model ensemble

November 2023

·

55 Reads

Atmospheric Science Letters

Atmospheric Science Letters

Using a large ensemble of initialised retrospective forecasts (hindcasts) from a seasonal prediction system, we explore various statistics relating to sudden stratospheric warmings (SSWs). Observations show that SSWs occur at a similar frequency during both El Niño and La Niña northern hemisphere winters. This is contrary to expectation, as the stronger stratospheric polar vortex associated with La Niña years might be expected to result in fewer of these extreme breakdowns. Here we show that this similar frequency may have occurred by chance due to the limited sample of years in the observational record. We also show that in these hindcasts, winters with two SSWs, a rare event in the observational record, on average have an increased surface impact. Multiple SSW events occur at a lower rate than expected if events were independent but somewhat surprisingly, our analysis also indicates a risk, albeit small, of winters with three or more SSWs, as yet an unseen event.


Characterising the observed SNAO, its relationship to jet variability and surface summer climate impacts
a MSLP regressed on to the 1st principal component associated with the 1st EOF of summer ERA5 MSLP calculated over the North Atlantic domain [90E-40W, 20-70 N] and accounting for 37% of the variance. Green boxes show regions used to calculate SNAO index²⁵. b the SNAO index correlated with the field of ERA5 upper level 300 hPa winds. Green box shows region used to calculate North Atlantic sector winds for cross-section plots in Figs. 3a, b and 5a. c the 1st EOF of upper level 300 hPa winds calculated over the extended domain shown and accounting for 20% of the variance. Contours in (b, c) are climatological 300 hPa zonal wind to highlight the position of the jets (contours plotted at 15 and 20 m/s). The SNAO index correlated with the ERA5 fields of summer precipitation (d) and linearly detrended 2 m air temperature (e). Stippling shows correlations significantly different from zero at the 90% confidence level according to a 2-sided Student’s t test.
Precursor SNAO signals in the late spring stratosphere
SNAO index correlated with the preceding May zonal mean zonal winds plotted as a latitude-height cross-section in ERA5 (a) and the model ensemble mean (b). Dashed horizontal line shows the approximate location of the tropopause, green boxes show the location of the box used to define the MPVI and stippling shows correlations significantly different from zero at the 95% confidence level according to a two-sided Student’s t test. c Histogram showing the distribution of model member correlations between the MPVI and SNAO. Pink solid and dashed vertical lines show the mean of the model member correlations and the 5-95% interval respectively, red vertical line shows the ensemble mean MPVI vs SNAO correlation and the black vertical lines shows the equivalent observed estimates from ERA5 and NCEP-R2 reanalyses⁶². d The strength of the SNAO vs MPVI correlation (red curve) as a function of ensemble size, with the black horizontals line showing the equivalent observed ERA5 and NCEP-R2 values. The green star shows the strength of SNAO vs MPVI correlation using the NCAR-SMYLE²⁹ prediction system for comparison (see Methods).
Downward propagation of late spring stratospheric anomalies
MPVI correlated with the monthly zonal mean winds over the North Atlantic sector (50–65 N, green box in Fig. 1b) plotted as a latitude-height cross-section in ERA5 (a) and the model ensemble mean (b). Solid vertical lines indicate the start and end of boreal summer, dashed horizontal line shows the approximate location of the tropopause, green stars indicate the location of the MPVI. MPVI correlated with subsequent summer upper level 300 hPa zonal winds in ERA (c) and model ensemble mean (d), contours are climatological 300 hPa zonal wind to highlight the position of the jets (contours plotted at 15 and 20 m/s). Stippling in (a–d) shows correlations significantly different from zero at the 90% confidence level according to a two-sided Student’s t test. e Impact of the MPVI on summer upper level 300 hPa zonal winds as diagnosed by perturbation experiments that switch atmospheric initial states to separate atmosphere and ocean drivers (Methods). Differences (m/s) are plotted between the average of the three positive and three negative MPVI states and stippling shows locations where these are significant at the 95% confidence level.
Windows of opportunity for skilful SNAO prediction and spuriously weak model signals
Standardised timeseries of ERA5 SNAO and the model predicted MPVI (a) and the model predicted SNAO (b). In both panels, all years are shown in the background, whilst the 21 years where the model predicted MPVI is ‘active’ (>±1 m/s) are plotted in the foreground. c model SNAO skill in predicting ERA5 SNAO (red) and itself (i.e. perfect predictability, green) as a function of ensemble size during active MPVI years. The switch to a thicker green line indicates the ensemble size where the model skill in predicting the observed SNAO is significantly higher than the skill of the model predicting itself. d histogram showing the distribution of the strength of SNAO persistence through the summer measured by SNAO in June correlated with SNAO in July/August. Solid and dashed vertical lines show the model member mean average and 5–95% intervals, respectively, the solid red vertical line shows the ensemble mean persistence and the vertical black line shows the ERA5 SNAO persistence.
Long-lasting influence of a late winter SSW
All panels are composite differences of observed ERA5 data for the 12 years in which an SSW occurred between 1st February and 15th March against the remaining 32 years (see Methods). a the composite difference for monthly zonal mean winds (m/s) over the North Atlantic sector (50–65N, green box in Fig. 1b) plotted as a latitude-height cross-section in ERA5 from February through to September. Composite differences for ERA5 upper troposphere 300 hPa zonal winds (m/s) in March (b) and summer (c), contours are climatological 300 hPa zonal wind to highlight the position of the jets (contours plotted at 15 and 20 m/s). Composite differences for ERA5 detrended 2 m air temperature (K) in March (d) and summer (e), green boxes highlight North America and Northern Europe which both transition from significant cold anomalies in March to warm ones in summer. Stippling on all panels shows significant differences at the 90% confidence level as assessed using a one-sided Student’s t test.
Skilful predictions of the Summer North Atlantic Oscillation

November 2023

·

228 Reads

·

4 Citations

Communications Earth & Environment

The Summer North Atlantic Oscillation is the primary mode of atmospheric variability in the North Atlantic region and has a significant influence on regional European, North American and Asian summer climate. However, current dynamical seasonal prediction systems show no significant Summer North Atlantic Oscillation prediction skill, leaving society ill-prepared for extreme summers. Here we show an unexpected role for the stratosphere in driving the Summer North Atlantic Oscillation in both observations and climate prediction systems. The anomalous strength of the lower stratosphere polar vortex in late spring is found to propagate downwards and influence the Summer North Atlantic Oscillation. Windows of opportunity are identified for useful levels of Summer North Atlantic Oscillation prediction skill, both in the 50% of years when the late spring polar vortex is anomalously strong/weak and possibly earlier if a sudden stratospheric warming event occurs in late winter. However, we show that model dynamical signals are spuriously weak, requiring large ensembles to obtain robust signals and we identify a summer ‘signal-to-noise paradox’ as found in winter atmospheric circulation Our results open possibilities for a range of new summer climate services, including for agriculture, water management and health sectors.


Windows of opportunity for predicting seasonal climate extremes highlighted by the Pakistan floods of 2022

October 2023

·

485 Reads

·

5 Citations

Nature Communications

Skilful predictions of near-term climate extremes are key to a resilient society. However, standard methods of analysing seasonal forecasts are not optimised to identify the rarer and most impactful extremes. For example, standard tercile probability maps, used in real-time regional climate outlooks, failed to convey the extreme magnitude of summer 2022 Pakistan rainfall that was, in fact, widely predicted by seasonal forecasts. Here we argue that, in this case, a strong summer La Niña provided a window of opportunity to issue a much more confident forecast for extreme rainfall than average skill estimates would suggest. We explore ways of building forecast confidence via a physical understanding of dynamical mechanisms, perturbation experiments to isolate extreme drivers, and simple empirical relationships. We highlight the need for more detailed routine monitoring of forecasts, with improved tools, to identify regional climate extremes and hence utilise windows of opportunity to issue trustworthy and actionable early warnings.


Variability and predictability of length of day
a, Variations in the length of day (LOD) showing the prominent interannual variability of around 0.5 × 10⁻³ s in observations (black) and the first year of ensemble mean model predictions starting in November each year (red). b, Correlation of predicted seasonal length-of-day anomalies in the ensemble mean with length-of-day anomalies from single model ensemble members (black), with radio telescope observations of Earth’s rotation (blue) and with atmospheric reanalysis (red). The perfect model predictability (black) is smoother than the prediction skill against observations (red, blue) due to averaging of the correlations with each ensemble member in the model case. Note the non-monotonic variation with lead time and the peaks at leads of 3 and 15 months in winter. Statistical significance at the 95% level according to a one-sided t test for positive correlations is shown by the dotted line.
Observed and predicted AAM fluctuations
a,b, Latitude–time plots of zonally integrated AAM anomalies are shown for a 20-year sample period from observational analysis (a) and corresponding ensemble means for the first year of each forecast from 1980 to 2000 (b). Vertical lines represent November each year, when forecasts are initialized. Units are 10²⁴ kg m² s⁻¹ per 0.55° latitude band, and the mean seasonal cycle has been removed from each latitude. Observational analyses are taken from the ERA datasets35,36. c, The correlation skill of the ensemble mean predictions as a function of latitude and lead time (months) for the whole period (1960–2017) using running seasonal means and a 6° latitudinal averaging.
Mechanism of wave-driven atmospheric anomalies
a, Predicted anomalies in AAM (black, 10²⁴ kg m² s⁻¹) and wave-driven acceleration (blue, 10⁻⁴ m s⁻²) from all waves (solid) and stationary waves (dotted) as a function of latitude, for the difference between the highest and lowest predicted AAM years. Anomalies are plotted from late spring/early summer (April–June) when ENSO anomalies decline to zero and the direct effect of ENSO from the tropics is small. Note how the wave driving accelerates the flow on the poleward side of the AAM anomaly in each hemisphere and how the stationary wave component is small compared with the total, implying that transient waves supply most of the wave-driven body force. b–d, A schematic of the wave-driven poleward propagation process for a positive perturbation to the AAM and zonal winds. b, The climatological jets (dashed line), spectrum of transient waves (wavy lines), climatological eddy momentum flux convergence (red) and divergence (blue). c,d, Initial perturbed jets and anomalies in eddy momentum flux divergence (c) and as the jet perturbation migrates poleward (d). Note that the same mechanism operates if the sign of anomalies is reversed.
AAM fluctuations precede changes in extratropical climate
a,b, Correlation between the ensemble mean predicted AAM (from forecasts started in November) and the following observed winter NAO (a) and Pacific jet-stream winds (b). Forecasts were started in November, and the NAO and jet-stream winds are predicted at a lead time of 13 months for all years between 1960 and 2017, inclusive. The NAO is the two-point difference in sea-level pressure between the Azores and Iceland, and the jet-stream wind is the zonal mean wind at 300 hPa and 60° N averaged over the Pacific (150° E to 150° W). The correlation with the following winter NAO and winds is plotted at each latitude and for each month as the forecasts progress. Positive correlations indicate that AAM anomalies precede the same-sign NAO and winds in the following winter as expected. Note the poleward migration with lead time (months), consistent with predictability arising from the poleward-migrating AAM anomalies. Hatching shows regions where the correlation between AAM and NAO is significant at the 90% level according to a one-sided t test.
Long-range predictability of extratropical climate and the length of day

October 2022

·

340 Reads

·

5 Citations

Nature Geoscience

Angular momentum is fundamental to the structure and variability of the atmosphere and therefore has an important influence on regional weather and climate. Total atmospheric angular momentum is also directly related to the rotation rate of the Earth and, hence, the length of day. However, the long-range predictability of fluctuations in the length of the day and atmospheric angular momentum is unknown. Here we show that fluctuations in atmospheric angular momentum and the length of day are predictable out to more than a year ahead and that this provides an atmospheric source of long-range predictability for surface climate. Using ensemble forecasts from a dynamical climate model, we demonstrate long-range predictability of signals in the atmospheric angular momentum field that propagate slowly and coherently polewards due to wave–mean flow interaction within the atmosphere. These predictable signals are also shown to precede changes in extratropical climate via the North Atlantic Oscillation and the extratropical jet stream. These results extend the lead time for length-of-day predictions, provide a source of long-range predictability from within the atmosphere and provide a link between geodesy and climate prediction.


Predictability of European winter 2020/2021: Influence of a mid‐winter sudden stratospheric warming

August 2022

·

38 Reads

·

4 Citations

Atmospheric Science Letters

Atmospheric Science Letters

Abstract Boreal winter (December–February) 2020/2021 in the North Atlantic/European region was characterised by a negative North Atlantic Oscillation (NAO) index. Although this was captured within the ensemble spread of predictions from the Met Office Global Seasonal forecast system (GloSea5), with 17% of ensemble members predicting an NAO less than zero, the forecast ensemble mean was shifted towards a positive NAO phase. The observed monthly NAO anomalies were particularly negative in January and February, following an early January sudden stratospheric warming (SSW), and a prolonged period of Phase 6 or 7 of the Madden Julian Oscillation (MJO) in late January/early February. In contrast, predictions showed the expected teleconnection from the observed La Niña, with a positive NAO signal resulting from a weakening of the Aleutian Low leading to a reduction in tropospheric wave activity, an increase in polar vortex strength and a reduced chance of an SSW. Forecasts initialised later in the winter season successfully predicted the negative NAO in January and February once the SSW and MJO were within the medium range timescale. GloSea5 likely over‐predicted the strength of the La Niña which we estimate caused a small negative bias in the SSW probability. However, this error is smaller than the uncertainty in SSW probability from the finite forecast ensemble size, emphasising the need for large forecast ensembles. This case study also demonstrates the advantage of continuously updated lagged ensemble forecasts over a ‘burst’ ensemble started on a fixed date, since a change in forecast signal due to events within the season can be detected early and promptly communicated to users.


FIGURE 1 | Comparison of skewness in Met Office Hadley Centre models. Skewness in 3 PPE members (A-C) and observations (HadISST, 1873-2020) (D). Rectangular box shows skewness in the far eastern Pacific region (FEP, 110-80 • W, 5 • N-5 • S). Skewness in FEP for PPE members (blue), high, medium, and low resolution HadGEM-CG3.1 simulations (brown), and observations (black) (E). Large circles show skewness over the full period of each model simulation, vertical bars show the 16th and 84th percentiles (corresponding to 1 standard deviation from the mean) and small circles are the maximum and minimum of skewness for 50-year sliding windows.
FIGURE 2 | Relationship between skewness and equatorial Pacific annual mean state. Spatial plot of grid point correlation of SST (A) zonal wind stress (C) and depth of 20 • isotherm (E) with FEP skewness. Scatter diagram of FEP skewness against annual mean Niño3 SST (B), central Pacific wind stress (D) and depth of 20 • isotherm for region Niño4 (F) for PPE members (small blue circles, with SK+ magenta, STD red, SK-cyan) and reanalysis data (large black circle). Regions for FEP (110-80 • W, 5 • N-5 • S), Niño3 (150-90 • W, 5 • N-5 • S), central Pacific (160 • E-80 • W, 5 • N-5 • S) and Niño4 (160 • E-150 • W, 5N-5S) are shown by boxes in the left-hand panel.
FIGURE 3 | Relationship between skewness and westerly wind burst activity. Scatter diagram of FEP SST skewness against annual mean wind burst activity (A) and against the El Niño minus La Niña difference in wind burst activity during the January to November period prior to ENSO events (B) for PPE members (small blue circles, with SK+ magenta, STD red, SK-cyan) and reanalysis data (large black circle). Horizontal bars show the standard error.
FIGURE 4 | Warming due to non-linear advection. Non-linear advection (K month −1 ) for combined ocean analyses (A) and combined PPE members (B). Box shows the averaging region for the non-linear advection. Scatter diagram of FEP skewness against non-linear advection for PPE ensemble members (small blue circles, with SK+ magenta, STD red, SK-cyan) and ocean reanalyses (large black circles) (C). Non-linear advection terms during growth period of El Niño (red) and La Niña (blue) for combined ocean reanalyses (D) combined PPE (E). Combined results are the average over the respective datasets.The growth period is defined here as the average of months April to November prior to the event, with the DJF El Niño/La Niña events identified using a ±0.8 • C threshold. Data in (C) are area-averages taken over the box shown in the upper panels (180 • -100 • W,1 • S-1 • N, 50-150 m). The combined ocean analysis data (A) have been smoothed. Vertical bars in (D,E) show the 95% confidence interval.
FIGURE 5 | Decomposition of zonal component of non-linear advection. Zonal component of non-linear advection (K month −1 ) for combined ocean reanalysis (A) and combined PPE (B), zonal gradient of temperature anomaly (K m −1 ) for combined ocean reanalysis (C) and combined PPE (D), and zonal current anomaly (ms −1 ) for combined ocean reanalysis (E) and combined PPE (F). Sections are for 1 • N-1 • S, for the growth period of ENSO, El Niño plus La Niña for (A,B) and El Niño minus La Niña for (C-F). The combined ocean analysis data have been smoothed.
ENSO Amplitude Asymmetry in Met Office Hadley Centre Climate Models

December 2021

·

114 Reads

·

2 Citations

Frontiers in Climate

Long climate simulations with the Met Office Hadley Centre General Circulation Model show weak El Niño-Southern Oscillation (ENSO) amplitude asymmetry between El Niño and La Niña phases compared with observations. This lack of asymmetry is explored through the framework of a perturbed parameter experiment. Two key hypotheses for the lack of asymmetry are tested. First, the possibility that westerly wind burst activity is biased is explored. It is found that the observed difference in wind burst activity during El Niño and La Niña tends to be underestimated by the model. Secondly, the warming due to subsurface non-linear advection is examined. While the model exhibits non-linear dynamic warming during both La Niña and El Niño, and thus a contribution to ENSO asymmetry, it is shown to be consistently underestimated in comparison with ocean reanalyses. The non-linear zonal advection term contributes most to the deficiency and the simulation of the anomalous zonal currents may be playing a key role in its underestimation. Compared with the ocean reanalyses, the anomalous zonal currents associated with ENSO are too weak in the vicinity of the equatorial undercurrent and the surface wind driven zonal currents extend too deep.


Evolutions of the 3-month running averaged Niño3.4 index (units: °C, colored thin curves) and their composites (red thick curve) for historical (a) El Niño and (b) La Niña events. Evolution starts from January of the event-developing year to December of next year (marked with +). (c) Probability histogram of ENSO peaking time (red bars, units: 1) and monthly variance of the Niño3.4 index evolution (green curve, units: °C²), both based on data from January 1900 to December 2019 (in (a) and (b), only events after the 1950s are shown). Red text on the upper-right indicates the count of events.
Probability histogram of ENSO peak time (red bars, left axis, units: 1) and monthly variance evolution of the Niño3.4 index (green curve, right axis, units: 1) for (a) CMIP5 and (b) CMIP6 MMEs. The Niño3.4 index used for MME was pre-normalized by each model’s standard deviation. Red text on the upper-right displays the count of events. Scatterplots of the locking-month (horizontal axis) and the sharpness of ENSO phase-locking (vertical axis, units: month²) for (c) CMIP5 and (d) CMIP6 models. Blue (green) numbers represent the corresponding CMIP5 (CMIP6) models listed in table 1. The red solid circle indicates the observational situation, and the blue (green) one indicates the CMIP5 (CMIP6) MME. The yellow area marks models reaching the level of relatively good simulation (simulating a locking-month during November-January and its sharpness with distance variance lower than 8.0).
Metrics of the locking-month of ENSO peak for CMIP5 (blue solid circles) and CMIP6 (green solid circles) models and their corresponding peak time of seasonal variation of SSTA amplitude (red oblique-crosses). The red triangle located on ‘Ob’ indicates the observational situation, and the blue (green) triangle located on ‘E5’ (‘E6’) indicates the CMIP5 (CMIP6) MME. The purple hollow circles and oblique-crosses indicate the second peak of the histogram and SSTA standard deviations in some models. Numbers represent the corresponding CMIP5 (CMIP6) models listed in table 1.
(a) Scatterplot of the intensity of seasonal modulation of ENSO’s instabilities (horizontal axis, units: 1) and the distance variance of phase-locking sharpness (vertical axis, units: month²). Blue (green) numbers represent the corresponding CMIP5 (CMIP6) models listed in table 1. The red solid circle indicates the observational situation, and the blue (green) one indicates the CMIP5 (CMIP6) MME. Hollow circles mark that models display a strong semi-annual variation of ENSO’s instabilities. (b) is similar to (a), but for scatters of the noise’s relative intensity (horizontal axis, units: 1) and the distance variance of phase-locking sharpness (vertical axis, units: month²), and models showing a strong semi-annual variation of ENSO’s instabilities are removed.
ENSO phase-locking behavior in climate models: from CMIP5 to CMIP6

April 2021

·

401 Reads

·

14 Citations

The phase-locking behavior of El Niño-Southern Oscillation (ENSO) in models from Coupled Model Intercomparison Project (CMIP) phase 5 to phase 6 is assessed in terms of the locking-month of ENSO peak and the sharpness of locking tendency. Overall, a robust improvement exists in CMIP6. Compared to CMIP5, more CMIP6 models truly reproduce the locking-month in November-January. Meanwhile, the sharpness of phase-locking in CMIP6 models also improves, though most of them are still far from the observations. The locking-month is verified to be highly corresponding to the phase of seasonal modulation of ENSO’s instabilities. The sharpness is mainly controlled by the intensity of this modulation and noise. Compared to CMIP5, CMIP6 models generally simulate these affecting factors better. Besides, models displaying an exaggerated semi-annual variation of ENSO’s instabilities simulate the ENSO phase-locking relative-poorly, and these models show no reduction from CMIP5 to CMIP6.


Predictability of European Winters 2017/2018 and 2018/2019: Contrasting influences from the Tropics and stratosphere

September 2020

·

123 Reads

·

21 Citations

Atmospheric Science Letters

Atmospheric Science Letters

The European winters of 2017–18 and 2018–19 were not climatically extreme, but both winters had a major sudden stratospheric warming (SSW). In February 2018, an SSW led to an intense cold outbreak across Europe and further spells of cold weather in March. The SSW of January 2019, although well predicted and expected to increase the chance of a cold end to winter, apparently produced little impact. In this study, we examine the performance of the Met Office seasonal prediction system in these winters, and the influences that led to these outcomes. To achieve this latter objective, sets of numerical experiments are performed in which the tropical troposphere and the extratropical stratosphere are relaxed towards their observed state, allowing the influence of each on the North Atlantic‐European atmospheric circulation to be identified. Using these experiments, we show that the SSWs had similar impacts in each case, creating a signal of easterly surface wind anomalies in the weeks following the event. In contrast, tropical influences were opposite in the two winters, acting to strengthen the easterly signal at the end of February 2018 and opposing it in January 2019. The different apparent responses to the two events therefore came about largely through tropical tropospheric variability. Furthermore, we highlight the importance of a very strong cycle of the Madden‐Julian Oscillation (MJO) in late January and early February 2018 as an important driver for the February 2018 SSW. MJO teleconnections appear to have been critical in creating the large mid‐latitude wave 2 amplitude that has been identified as the immediate cause of this event.


Citations (49)


... The wave source in subpolar North Atlantic can be resultant from active air-sea interaction (Nie et al 2019), and thus the status of spring and early-summer SST and atmospheric circulation anomalies are important for the initialization of early-summer heat extremes. For example, a negative North Atlantic Oscillation response to the sudden stratospheric warming in spring may reduce heat loss from the ocean and then lead to persistent warm SST anomalies over subpolar North Atlantic (Dunstone et al 2023). The Eurasian jet stream is determined by both transient eddy forcing and meridional thermal contrast (Messori et al 2021), and may manifest different characteristics during different periods of the whole summer. ...

Reference:

Eurasian mid-latitude jet stream bridges an Atlantic to Asia summer teleconnection in heat extremes
Skilful predictions of the Summer North Atlantic Oscillation

Communications Earth & Environment

... It stands out as the most dominant interannual mode within the tropical-coupled ocean-atmosphere system (Mahala et al. 2015). These ENSO events typically have a duration of approximately 12-18 months and recur every 2-7 years, displaying considerable variability in their strengths (Dunstone et al. 2023). Throughout ENSO events, the atmospheric response to sea surface temperature (SST) anomalies in the equatorial Pacific disrupts the water circulation, exerting a profound influence on both oceanic and atmospheric conditions globally (Mahala et al. 2015). ...

Windows of opportunity for predicting seasonal climate extremes highlighted by the Pakistan floods of 2022

Nature Communications

... Nevertheless, their utilization has the potential to enhance the accuracy of EOP predictions for a maximum forecast horizon of about two weeks. Future research aiming to improve EOP predictions should encompass the development of new EAM data and longer-term EAM forecasts, by making use of even longer prediction runs that are now being performed by various numerical weather prediction centers (Scaife et al. 2022). ...

Long-range predictability of extratropical climate and the length of day

Nature Geoscience

... Prominent global drivers of the anomalous climatic conditions at the time included the La Niña, which was present for the second winter in a row. La Niña winters tend to have blocked Atlantic teleconnection patterns in early winter (November-December) and mobile westerly patterns over the North Atlantic and northern Europe in late winter (e.g., Ayarzaguena et al., 2018;Dunstone et al., 2018;Fereday et al., 2008;King et al., 2018;Lockwood et al., 2022;Molteni & Brookshaw, 2023). In addition, there was also an easterly phase of the equatorial stratospheric Quasi-Biennial Oscillation which tends to weaken the strength of the stratospheric polar vortex (SPV) (e.g. ...

Predictability of European winter 2020/2021: Influence of a mid‐winter sudden stratospheric warming
Atmospheric Science Letters

Atmospheric Science Letters

... The east Pacific "cold tongue" is often too cold in models (Li & Xie, 2014), and models typically exhibit the "double intertropical convergence zone" phenomenon, whereby the observed zonal band of precipitation known as the intertropical convergence zone is too far north, a second band forms below the Equator, and precipitation on the Equator itself is underestimated (Lin, 2007). The observed asymmetry of the distribution of eastern Pacific SST anomalies is also not generally captured by models (Ineson et al., 2021;Zhang & Sun, 2014). This affects the simulation of impacts due to strong El Niño events, as well as whether the model produces such events at all. ...

ENSO Amplitude Asymmetry in Met Office Hadley Centre Climate Models

Frontiers in Climate

... The 500-year integration period is chosen in the present study. Besides, based on the previous result of the model evaluation, we chose the CCSM4 climate model, which owns superior simulation skills for two types of El Niño events as well as the interaction process between tropical Pacific and other basins (Freund et al. 2020;Ham and Kug, 2015;Hou et al. 2022;Kim and Yu, 2012;Kucharski et al. 2015;Liu et al. 2021;Planton et al. 2021). The 500 years of sea surface temperature (SST), precipitation flux, and zonal and meridional wind monthly components derived from CCSM4 output are used. ...

ENSO phase-locking behavior in climate models: from CMIP5 to CMIP6
Environmental Research Communications

Environmental Research Communications

... A prime example consists of the weak vortex conditions in a sudden stratospheric warming (SSW), which is followed by a southward shift and weakened tropospheric jet stream in the mid-latitude region (Baldwin et al., 2021;Kidston et al., 2015). However, it is worth noting that certain prominent SSW events elicit varying tropospheric responses Knight et al., 2021), and the relationship between the tropospheric events and stratospheric variability is not always consistent. Although the impacts of the stratosphere on lower atmospheric circulation were once debated (Plumb & Semeniuk, 2003), numerical experiments have demonstrated that the perturbed stratosphere can reproduce surface climatic responses with observations (Douville, 2009;Hitchcock & Haynes, 2016;White et al., 2020). ...

Predictability of European Winters 2017/2018 and 2018/2019: Contrasting influences from the Tropics and stratosphere
Atmospheric Science Letters

Atmospheric Science Letters

... The El Niño/Southern Oscillation (ENSO)-the dominant mode of climate variability at multi-year time scales-influences global weather via atmospheric teleconnections (Lenssen et al., 2020;Mason & Goddard, 2001;Ropelewski & Halpert, 1986), and has well-known predictability at lead times of nine or fewer months Tippett et al., 2019;L'Heureux et al., 2020;Becker et al., 2022). Numerous forecast systems have shown small, but significant predictive skill at lead times beyond 9 months with dynamical (Dunstone et al., 2020;Gonzalez & Goddard, 2016) and statistical (Ding & Alexander, 2023;Ham et al., 2019;Wang et al., 2023) methods, but the sources of this skill are not firmly established. ...

Skilful interannual climate prediction from two large initialised model ensembles

... The most recent phase is CMIP6 (Bock and Lauer, 2023;Eyring et al., 2016Eyring et al., , 2019Goldenson et al., 2018;Stouffer et al., 2017). CMIP6 models have shown improved performance in simulating historical precipitation compared to their CMIP5 counterparts globally and for different areas worldwide (Lun et al., 2021;Menary et al., 2020;Swart et al., 2019;Zhu and Yang, 2020). These enhancements are primarily attributed to advancements in model physics, parameterizations and resolution, as well as improved representation of key climate processes such as aerosol-cloud interactions and land-atmosphere coupling (Kay et al., 2015). ...

Preindustrial Control Simulations With HadGEM3‐GC3.1 for CMIP6

... Li and Lau (2012) and Jiménez-Esteve and Domeisen (2018) find that the southward shifted Pacific jet stream increases the eastward propagation of transient eddies to the NAE region, which favors the negative phase of the NAO, whereas Drouard et al. (2015) show that circulation anomalies in the North Pacific influence Rossby wave breaking in the Atlantic sector, subsequently affecting the phase of the NAO. ENSO-related anomalies in the tropical and extratropical Atlantic and the Caribbean have also been shown to exert an influence (Hardiman et al., 2019;Herceg-Bulić et al., 2023;Toniazzo & Scaife, 2006). ...

The Impact of Strong El Niño and La Niña Events on the North Atlantic