Skip to main content
Log in

Stochastic investigation of daily air temperature extremes from a global ground station network

  • Original Paper
  • Published:
Stochastic Environmental Research and Risk Assessment Aims and scope Submit manuscript

Abstract

Near-surface air temperature is one of the most widely studied hydroclimatic variables, as both its regular and extremal behaviors are of paramount importance to human life. Following the global warming observed in the past decades and the advent of the anthropogenic climate change debate, interest in temperature’s variability and extremes has been rising. It has since become clear that it is imperative not only to identify the exact shape of the temperature’s distribution tails, but also to understand their temporal evolution. Here, we investigate the stochastic behavior of near-surface air temperature using the newly developed estimation tool of Knowable (K-)moments. K-moments, because of their property to substitute higher-order deviations from the mean with the distribution function, enable reliable estimation and an effective alternative to order statistics and, particularly for the outliers-prone distribution tails. We compile a large set of daily timeseries (30–200 years) of average, maximum and minimum air temperature, which we standardize with respect to the monthly variability of each record. Our focus is placed on the maximum and minimum temperatures, because they are more reliably measured than the average, yet very rarely analyzed in the literature. We examine segments of each timeseries using consecutive rolling 30-year periods, from which we extract extreme values corresponding to specific return period levels. Results suggest that the average and minimum temperature tend to increase, while overall the maximum temperature is slightly decreasing. Furthermore, we model the temperature timeseries as a filtered Hurst-Kolmogorov process and use Monte Carlo simulation to produce synthetic records with similar stochastic properties through the explicit Symmetric Moving Average scheme. We subsequently evaluate how the patterns observed in the longest records can be reproduced by the synthetic series.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Source: Jones et al. (2016)

Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Code availability

The scripts and functions used, all of which were implemented in Matlab can be downloaded from www.itia.ntua.gr/2079/. Also, a readme file, in txt format, on the same repository contains explanatory information on the operations each code script performs.

References

  • Aram F, García EH, Solgi E, Mansournia S (2019) Urban green space cooling effect in cities. Heliyon 5(4):e01339

    Article  Google Scholar 

  • Batchelor GK, Townsend AA (1949) The nature of turbulent motion at large wave-numbers. Proc R Soc Lond Ser A Math Phys Sci 199(1057):238–255

    Google Scholar 

  • Bernatzky A (1982) The contribution of tress and green spaces to a town climate. Energy Build 5(1):1–10

    Article  Google Scholar 

  • Braganza K, Karoly DJ, Arblaster JM (2004) Diurnal temperature range as an index of global climate change during the twentieth century. Geophys Res Lett. https://doi.org/10.1029/2004GL019998

    Article  Google Scholar 

  • Brown PT, Li W, Cordero EC, Mauget SA (2015) Comparing the model-simulated global warming signal to observations using empirical estimates of unforced noise. Sci Rep 5:9957

    Article  CAS  Google Scholar 

  • Cavanaugh NR, Shen SS (2014) Northern Hemisphere climatology and trends of statistical moments documented from GHCN-daily surface air temperature station data from 1950 to 2010. J Clim 27(14):5396–5410

    Article  Google Scholar 

  • Coumou D, Rahmstorf S (2012) A decade of weather extremes. Nat Clim Change 2(7):491–496. https://doi.org/10.1038/nclimate1452

    Article  Google Scholar 

  • Cronin TM (2009) Paleoclimates: understanding climate change past and present. Columbia University Press, New York

    Google Scholar 

  • Dimitriadis P (2017) Hurst-Kolmogorov dynamics in hydrometeorological processes and in the microscale of turbulence. Ph.D. Thesis, Department of Water Resources and Environmental Engineering—National Technical University of Athens, Athens, Greece

  • Dimitriadis P, Koutsoyiannis D (2015) Climacogram versus autocovariance and power spectrum in stochastic modelling for Markovian and Hurst-Kolmogorov processes. Stoch Env Res Risk Assess 29(6):1649–1669

    Article  Google Scholar 

  • Dimitriadis P, Koutsoyiannis D (2018) Stochastic synthesis approximating any process dependence and distribution. Stoch Environ Res Risk Assess 32(6):1493–1515

    Article  Google Scholar 

  • Dimitriadis P, Koutsoyiannis D (2019) The mode of the climacogram estimator for a Gaussian Hurst-Kolmogorov process. J Hydroinform. https://doi.org/10.2166/hydro.2019.038

    Article  Google Scholar 

  • Dimitriadis P, Tzouka K, Koutsoyiannis D, Tyralis H, Kalamioti A, Lerias E, Voudouris P (2019) Stochastic investigation of long-term persistence in two-dimensional images of rocks. Spat Stat 29:177–191. https://doi.org/10.1016/j.spasta.2018.11.002

    Article  Google Scholar 

  • Dittus AJ, Karoly DJ, Lewis SC, Alexander LV (2015) A multiregion assessment of observed changes in the areal extent of temperature and precipitation extremes. J Clim 28(23):9206–9220

    Article  Google Scholar 

  • Easterling DR, Horton B, Jones PD, Peterson TC, Karl TR, Parker DE, Folland CK (1997) Maximum and minimum temperature trends for the globe. Science 277(5324):364–367

    Article  CAS  Google Scholar 

  • Efstratiadis A, Dialynas YG, Kozanis S, Koutsoyiannis D (2014) A multivariate stochastic model for the generation of synthetic time series at multiple time scales reproducing long-term persistence. Environ Model Softw 62:139–152

    Article  Google Scholar 

  • Geiger R (1954) Klassifikation der klimate nach W. Köppen. Landolt-Börnstein–Zahlenwerte und Funktionen aus Physik. Chem Astron Geophys Tech 3:603–607

    Google Scholar 

  • Gneiting T, Schlather M (2004) Stochastic models that separate fractal dimension and the Hurst effect. SIAM Rev 46:269–282. https://doi.org/10.1137/s0036144501394387

    Article  Google Scholar 

  • Handmer J, Honda Y, Kundzewicz ZW, Arnell N, Benito G, Hatfield J, Takahashi K (2012) Changes in impacts of climate extremes: human systems and ecosystems. In: Managing the risks of extreme events and disasters to advance climate change adaptation special report of the intergovernmental panel on climate change (pp. 231–290). Intergovernmental Panel on Climate Change

  • Hasselmann K (1976) Stochastic climate models part I. Theory tellus 28(6):473–485

    Google Scholar 

  • Hirschi JJM, Sinha B (2007) Negative NAO and cold Eurasian winters: how exceptional was the winter of 1962/1963? Weather 62(2):43–48

    Article  Google Scholar 

  • Hurst HE (1951) Long-term storage capacity of reservoirs. Trans Amer Soc Civil Eng 116:770–799

    Article  Google Scholar 

  • Iliopoulou T, Koutsoyiannis D (2019) Revealing hidden persistence in maximum rainfall records. Hydrol Sci J 64(14):1673–1689. https://doi.org/10.1080/02626667.2019.1657578

    Article  Google Scholar 

  • IPCC (2014) Climate change 2013: the physical science basis: working group I contribution to the fifth assessment report of the intergovernmental panel on climate change. Cambridge University Press, Cambridge. https://doi.org/10.1017/CBO9781107415324

    Book  Google Scholar 

  • IPCC (2018) Summary for Policymakers. In: Global warming of 1.5 °C. An IPCC Special Report on the impacts of global warming of 1.5 °C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty. World Meteorological Organization, Geneva

  • Jones PD, Lister DH, Osborn TJ, Harpham C, Salmon M, Morice CP (2012) Hemispheric and large-scale land-surface air temperature variations: an extensive revision and an update to 2010. J Geophys Res Atmos. https://doi.org/10.1029/2011JD017139

    Article  Google Scholar 

  • Jones PD, Parker DE, Osborn TJ, Briffa KR (2016) Global and hemispheric temperature anomalies: land and marine instrumental records (1850–2015). Environmental system science data infrastructure for a virtual ecosystem; Carbon Dioxide Information Analysis Center (CDIAC), Oak Ridge National Laboratory, Oak Ridge, TN (USA). 10.3334/CDIAC/cli.002

  • Khasanov BF (2013) Severe winter rings of oak trees (Quercus robur L.) from Central European Russia. Int J Biometeorol 57(6):835–843

    Article  CAS  Google Scholar 

  • Kolmogorov AN (1940) Wienersche spiralen und einige andere interessante kurven in hilbertscen raum, cr (doklady). Acad Sci URSS (NS) 26:115–118

    Google Scholar 

  • Koutsoyiannis D (2000) A generalized mathematical framework for stochastic simulation and forecast of hydrologic time series. Water Resour Res 36(6):1519–1533

    Article  Google Scholar 

  • Koutsoyiannis D (2002) The Hurst phenomenon and fractional Gaussian noise made easy. Hydrol Sci J 47(4):573–595

    Article  Google Scholar 

  • Koutsoyiannis D (2010) HESS opinions a random walk on water. Hydrol Earth Syst Sci 14(3):585–601

    Article  Google Scholar 

  • Koutsoyiannis D (2011) Hurst-Kolmogorov dynamics and Uncertainty 1. JAWRA J Am Water Resour Assoc 47(3):481–495

    Article  Google Scholar 

  • Koutsoyiannis D (2014) Entropy: from thermodynamics to hydrology. Entropy 16(3):1287–1314

    Article  Google Scholar 

  • Koutsoyiannis D (2016) Generic and parsimonious stochastic modelling for hydrology and beyond. Hydrol Sci J 61(2):225–244

    Article  Google Scholar 

  • Koutsoyiannis D (2017) Entropy production in stochastics. Entropy 19(11):581

    Article  Google Scholar 

  • Koutsoyiannis D (2019a) Knowable moments for high-order stochastic characterization and modelling of hydrological processes. Hydrol Sci J 64(1):19–33

    Article  Google Scholar 

  • Koutsoyiannis D (2019b) Advances in stochastics of hydroclimatic extremes. Presentation. Conference: Giornata di studio in memoria di Baldassare Bacchi, University of Brescia, Italy. doi: https://doi.org/10.13140/RG.2.2.30655.05282/1

  • Koutsoyiannis D (2020) Stochastics of hydroclimatic extremes. National Technical University of Athens. http://itia.ntua.gr/2000/. Access Date 20 Dec 2020

  • Koutsoyiannis D, Dimitriadis P, Lombardo F, Stevens S (2018) From fractals to stochastics: seeking theoretical consistency in analysis of geophysical data. In: Tsonis AA (ed) Advances in Nonlinear Geosciences. Springer, Cham, pp 237–278. https://doi.org/10.1007/978-3-319-58895-7_14

    Chapter  Google Scholar 

  • Mandelbrot BB, Van Ness JW (1968) Fractional Brownian motions, fractional noises and applications. SIAM Rev 10(4):422–437

    Article  Google Scholar 

  • Masson-Delmotte TWV, Zhai P, Pörtner HO, Roberts D, Skea J, Shukla PR, Connors S (2018) IPCC, 2018: Summary for policymakers. In: Global warming of 1.5 °C. An IPCC special report on the impacts of global warming of 1.5 °C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global. World Meteorological Organization, Geneva, Technical Report

  • Menne MJ, Durre I, Korzeniewski B, McNeal S, Thomas K, Yin X, Houston TG (2012) Global historical climatology network-daily (GHCN-Daily), Version 3. NOAA National Climatic Data Center, 10, V5D21VHZ. Access Date: 15 Apr 2019

  • Montanari A, Rosso R, Taqqu MS (1997) Fractionally differenced ARIMA models applied to hydrologic time series: Identification, estimation, and simulation. Water Resour Res 33(5):1035–1044

    Article  Google Scholar 

  • O’Connell PE, Koutsoyiannis D, Lins HF, Markonis Y, Montanari A, Cohn T (2016) The scientific legacy of Harold Edwin Hurst. Hydrol Sci J 61:1571–1590. https://doi.org/10.1080/02626667.2015.1125998

    Article  Google Scholar 

  • Papoulis A (1990) Probability and statistics, vol 2. Prentice-Hall, Englewood Cliffs

    Google Scholar 

  • Peterson TC, Gallo KP, Lawrimore J, Owen TW, Huang A, McKittrick DA (1999) Global rural temperature trends. Geophys Res Lett 26(3):329–332

    Article  Google Scholar 

  • Portmann RW, Solomon S, Hegerl GC (2009) Spatial and seasonal patterns in climate change, temperatures, and precipitation across the United States. Proc Natl Acad Sci 106(18):7324–7329

    Article  CAS  Google Scholar 

  • Rubel F, Kottek M (2010) Observed and projected climate shifts 1901–2100 depicted by world maps of the Köppen-Geiger climate classification. Meteorol Z 19(2):135–141

    Article  Google Scholar 

  • Sargentis G-F, Dimitriadis P, Ioannidis R, Iliopoulou T, Koutsoyiannis D (2019) Stochastic evaluation of landscapes transformed by renewable energy installations and civil works. Energies 12:2817. https://doi.org/10.3390/en12142817

    Article  Google Scholar 

  • Sargentis G-F, Dimitriadis P, Koutsoyiannis D (2020) Aesthetical issues of Leonardo Da Vinci’s and Pablo Picasso’s paintings with stochastic evaluation. Heritage 3:283–305. https://doi.org/10.3390/heritage3020017

    Article  Google Scholar 

  • Sigourou S, Dimitriadis P, Iliopoulou T, Ioannidis R, Skopeliti A, Sakellari K, Koutsoyiannis D (2018) Statistical and stochastic comparison of climate change versus urbanization. In: EGU general assembly conference abstracts, Vol. 20, p. 18608

  • Sun X, Ren G, Xu W, Li Q, Ren Y (2017) Global land-surface air temperature change based on the new CMA GLSAT data set. Sci Bull 62(4):236–238

    Article  Google Scholar 

  • Trenberth KE, Jones PD, Ambenje P, Bojariu R, Easterling D, Klein Tank A, Parker D, Rahimzadeh F, Renwick JA, Rusticucci M, Soden B, Zhai P (2007) Observations: surface and atmospheric climate change, Chapter 3. Climate change, 235–336. Available at http://www.ipcc.ch/pdf/assessment-report/ar4/wg1/ar4-wg1-chapter3.pdf

Download references

Acknowledgements

We are grateful to the Editor in Chief George Christakos, the anonymous Associate Editor and the two anonymous reviewers for their efforts, useful comments and suggestions that helped us improve the paper.

Funding

No funds have been available for this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Konstantinos-Georgios Glynis.

Ethics declarations

Conflict of interest

All authors certify that they have no affiliations with or involvement in any organization or entity with any financial interest or non-financial interest in the subject matter or materials discussed in this manuscript.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file1 (36 KB)

Appendices

Appendix 1

To facilitate the understanding of the theory behind K-moments, we explain some basic notions of statistics in this appendix.

Let \(\underline {x}\) be a stochastic variable and \(\underline {x}_{1}\), \(\underline {x}_{2}\), …, \(\underline {x}_{p}\) be copies of it, independent and identically distributed, forming a sample. The maximum of all, which is identical to the pth order stochastic, is by definition:

$$\underline {x}_{\left( p \right)} := {\text{max}}\left( {\underline {x}_{1} , \underline {x}_{2} , \ldots , \underline {x}_{p} } \right)$$
(19)

It is readily obtained that if \(F\left( x \right)\) is the distribution function of \(\underline {x}\) and \(f\left( x \right)\) its probability density function, then those of \(\underline {x}_{\left( p \right)}\) are distributed by:

$$F^{\left( p \right)} \left( x \right) = \left( {F\left( x \right)} \right)^{p} ,\,\,f^{\left( p \right)} \left( x \right) = pf\left( x \right)\left( {F\left( {\underline {x} } \right)} \right)^{p - 1}$$
(20)

where the former is the product of \(p\) instances of \(F\left( x \right)\) (justified by the independent and identically distributed assumption), while the latter is the derivative of \(F^{\left( p \right)} \left( x \right)\) with respect to \(x\). The expected maximum order of \(p\) of \(\underline {x}\), i.e. the expected value of \(\underline {x}_{\left( p \right)}\), is therefore:

$${\text{E}}\left[ {\underline {x}_{\left( p \right)} } \right] = {\text{E}}\left[ {{\text{max}}\left( {\underline {x}_{1} , \underline {x}_{2} , \ldots ,\underline {x}_{p} } \right)} \right] = p{\text{E}}\left[ {\left( {F\left( {\underline {x} } \right)} \right)^{p - 1} \underline {x} } \right]$$
(21)

It is worth to stress that the variables \(\underline {x}_{1}\)., \(\underline {x}_{2}\), …, \(\underline {x}_{p}\) considered here, are not meant in temporal succession and, in this respect, do not form a stochastic process, but are rather regarded to be an ensemble of copies of \(\underline {x}\). In other words, the possible dependence in time of a stochastic process is not considered to be prerequisite for the application.

In geophysical processes, it is justifiable to assume that the variance \(\mu_{2} \equiv \sigma^{2}\) is finite, because an infinite variance would translate to an infinite amount of eney to materialize, which is absurd. However, high-order classical moments \(\mu_{p}\) diverge to infinity beyond a certain \(p\) (i.e., in heavy-tailed distributions). That is not the case for the K-moments, here a significant part of the moment is calculated using the always finite distributi function (Koutsoyiannis 2019a), which is the reason from which their knowability stems.

To derive knowable moments for high orders \(p\), in the expectation defining the pth moment, we raise \(\left( {\underline {x} - \mu } \right)\). to a low power \(q < p\) and for the remaining \(\left( {p - q} \right)\) multiplicative terms, we replace \(\left( {\underline {x} - \mu } \right)\) with \(\left( {2F\left( {\underline {x} } \right) - 1} \right),\) where \(F\left( x \right)\) is the distribution function. This leads to the following definition of central K-moment of order \(\left( {p,q} \right)\) (Koutsoyiannis 2019a):

$$K_{pq} : = \left( {p - q + 1} \right){\text{E}}\left[ {\left( {2F\left( {\underline {x} } \right) - 1} \right)^{p - q} \left( {\underline {x} - \mu } \right)^{q} } \right],\quad p \ge q$$
(22)

(Likewise, the non-central K-moment of order \(\left( {p,q} \right)\) is defined (Koutsoyiannis 2019a):

$$K^{\prime}_{pq} : = \left( {p - q + 1} \right){\text{E}}\left[ {\left( {F\left( {\underline {x} } \right)} \right)^{p - q} \underline {x}^{q} } \right], \quad p \ge q$$
(23)

The quantities \(\left( {F\left( {\underline {x} } \right)} \right)^{p - q}\) and \(\left( {2F\left( {\underline {x} } \right) - 1} \right)^{p - q}\) are estimated from a sample, without the use of powers of \(\underline {x}\), thus making the estimation more reliable. Specifically, for the ith element of a sample \(x_{\left( i \right)}\) of size \(n\), sorted in ascending order, \(F\left( {x_{\left( i \right)} } \right)\) and \(\left( {2F\left( {x_{\left( i \right)} } \right) - 1} \right)\) are estimated as:

$$\hat{F}\left( {x_{\left( i \right)} } \right) = \frac{i - 1}{{n - 1}}, \ldots 2\hat{F}\left( {x_{\left( i \right)} } \right) - 1 = \frac{2i - n - 1}{{n - 1}}$$
(24)

taking values in [0,1] and [− 1,1], respectively, irrespective of the values \(x_{\left( i \right)}\).. Hence, the estimators of K-moments are:

$$\hat{K}^{\prime}_{pq} = \frac{p - q + 1}{n}\mathop \sum \limits_{i = 1}^{n} \left( {\frac{i - 1}{{n - 1}}} \right)^{p - q} \underline {x}_{\left( i \right)}^{q}$$
(25)
$$\hat{K}_{pq} = \frac{p - q + 1}{n}\mathop \sum \limits_{i = 1}^{n} \left( {\frac{2i - n - 1}{{n - 1}}} \right)^{p - q} (\underline {x}_{\left( i \right)} - \hat{\mu })^{q}$$
(26)

The rationale of the definition is very relatively easy to grasp. Assuming that the distribution mean is close to the median, so that \(F\left( \mu \right) \approx {\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 2}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$2$}}\) (this is precisely true for a symmetric distribution), the quantity whose expectation is taken from the finition of the central K-moment of order \(\left( {p,q} \right)\) is: \(A\left( {\underline {x} } \right){ \sim } := \left( {2F\left( {\underline {x} } \right) - 1} \right)^{p - q} \left( {\underline {x} - \mu } \right)^{q}\) and its Taylor expansion is:

$$\begin{aligned} A\left( {\underline {x} } \right) &= \left( {2f\left( \mu \right)} \right)^{p - q} \left( {\underline {x} - \mu } \right)^{p} + \left( {p - q} \right)\left( {2f\left( \mu \right)} \right)^{p - q - 1} f^{\prime}\left( \mu \right)\left( {\underline {x} - \mu } \right)^{p + 1} \\&\quad+ O\left( {\left( {\underline {x} - \mu } \right)^{p + 2} } \right) \end{aligned}$$
(27)

where \(f\left( x \right)\) is the probability density function of \(\underline {x}\). Clearly then, \(K_{pq}\) depends on \(\mu_{p}\) as well as on classical moments of \(\underline {x}\) of order higher than \(p\). The independence of \(K_{pq}\) from classical moments of order smaller than \(p\) is the reason why it is a competent surrogate of the unknowable \(\mu_{p}\). In addition, as \(p\) becomes large, by virtue of the multiplicative term \(\left( {p - q + 1} \right)\) in the definition of K-moments, \(K_{pq}\) shares similar asymptotic properties with \(\hat{\mu }_{p}^{{{\raise0.7ex\hbox{$q$} \!\mathord{\left/ {\vphantom {q p}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$p$}}}}\) (the estimate, not the true \(\mu_{p}^{{{\raise0.7ex\hbox{$q$} \!\mathord{\left/ {\vphantom {q p}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$p$}}}}\)). To illustrate this for \(q = 1\) and for independent variables \(\underline {x}_{i}\), we consider the variable \(\underline {z}_{p} { \sim } = {\text{max}}_{1 \le i \le p} \underline {x}_{i}\) and denote \(f\left( \right)\) and \(h\left( \right)\). the probability densities of \(\underline {x}_{i}\). and \(\underline {z}_{i}\) respectively. Then (Papoulis 1990):

$$h\left( z \right) = pf\left( z \right)\left( {(F\left( z \right)} \right)^{p - 1}$$
(28)

and thus, by virtue of the definition of non-central K-moment of order \(\left( {p,q} \right)\):

$${\text{{\rm E}}}\left[ {\underline{z} _{p} } \right] = p{\text{E}}\left[ {\left( {F\left( {\underline{x} } \right)} \right)^{{p - 1}} \underline{x} } \right] = K'_{{p1}}$$
(29)

On the other hand, for positive \(\underline {x}\) and large \(p \to n\),

$$\begin{aligned} \left( {{\text{E}}\left[ {\underline{{\widehat{{\mu '}}}} _{P} } \right]} \right)^{{{\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 p}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$p$}}}} & = \left( {{\text{E}}\left[ {\left( {\frac{1}{n}\mathop \sum \limits_{{i = 1}}^{n} \underset{\raise0.3em\hbox{$\smash{\scriptscriptstyle-}$}}{x} _{i} ^{p} } \right)} \right]} \right)^{{{\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 p}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$p$}}}} \approx \left( {{\text{E}}\left[ {\left( {\frac{1}{n}{\text{max}}_{{1 \le i \le n}} \left( {x_{i}^{p} } \right)} \right)} \right]} \right)^{{{\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 p}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$p$}}}} \\ & \approx n^{{ - {\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 p}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$p$}}}} {\text{E}}\left[ {{\text{max}}_{{1 \le i \le n}} \left( {\underset{\raise0.3em\hbox{$\smash{\scriptscriptstyle-}$}}{x} _{i} } \right)} \right] \approx {\text{E}}\left[ {\underset{\raise0.3em\hbox{$\smash{\scriptscriptstyle-}$}}{z} _{n} } \right] \\ \end{aligned}$$
(30)

It is also worth noting that the multiplicative term \(\left( {p - q + 1} \right)\) in the definitions of central and non-central \(K_{pq}\) and \(K^{\prime}_{pq}\) makes K-moments generally increasing functions of \(p\).

Appendix 2

The Climacograms of the three parameters of the near-surface air temperature (average, maximum and minimum) are presented in the following figures. Note that the climacogram derived from the empirical data is depicted in blue color, while the climacogram of the synthetic data is in green color respectively. Solid lines represent the mean of each dataset (empirical and synthetic), while dashed lines represent the 5th and 95th percentile (90% confidence levels) of the respective distributions. The climacogram derived from the optimally fitted theoretical model is depicted in red colored solid line.

It is worth noting that the range between the 5th and 95th percentiles of the synthetic data in each of the three climacograms is narrower than the expected one from the respective empirical data. This is probably caused by the use of the same model (imposed by the same Hurst and Mandelbrot parameters) in the production of the synthetic timeseries for each of the three parameters of near-surface air temperature (Figs. 14, 15, 16)

Fig. 14
figure 14

Climacogram of the average air temperature

Fig. 15
figure 15

Climacogram of the maximum air temperature

Fig. 16
figure 16

Climacogram of the minimum air temperature

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Glynis, KG., Iliopoulou, T., Dimitriadis, P. et al. Stochastic investigation of daily air temperature extremes from a global ground station network. Stoch Environ Res Risk Assess 35, 1585–1603 (2021). https://doi.org/10.1007/s00477-021-02002-3

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00477-021-02002-3

Keywords

Navigation