Advanced Search
Article Contents

The Reliability of Global and Hemispheric Surface Temperature Records


doi: 10.1007/s00376-015-5194-4

  • The purpose of this review article is to discuss the development and associated estimation of uncertainties in the global and hemispheric surface temperature records. The review begins by detailing the groups that produce surface temperature datasets. After discussing the reasons for similarities and differences between the various products, the main issues that must be addressed when deriving accurate estimates, particularly for hemispheric and global averages, are then considered. These issues are discussed in the order of their importance for temperature records at these spatial scales: biases in SST data, particularly before the 1940s; the exposure of land-based thermometers before the development of louvred screens in the late 19th century; and urbanization effects in some regions in recent decades. The homogeneity of land-based records is also discussed; however, at these large scales it is relatively unimportant. The article concludes by illustrating hemispheric and global temperature records from the four groups that produce series in near-real time.
  • 加载中
  • Arnfield A. J.,2003: Two decades of urban climate research: a review of turbulence, exchanges of energy and water, and the urban heat island. Inter. J. Climatol., 23, 1-26.10.1002/joc.859fbef57281fe086d55a541959f110c95fhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.859%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/joc.859/fullAbstract Progress in urban climatology over the two decades since the first publication of the International Journal of Climatology is reviewed. It is emphasized that urban climatology during this period has benefited from conceptual advances made in microclimatology and boundary-layer climatology in general. The role of scale, heterogeneity, dynamic source areas for turbulent fluxes and the complexity introduced by the roughness sublayer over the tall, rigid roughness elements of cities is described. The diversity of urban heat islands, depending on the medium sensed and the sensing technique, is explained. The review focuses on two areas within urban climatology. First, it assesses advances in the study of selected urban climatic processes relating to urban atmospheric turbulence (including surface roughness) and exchange processes for energy and water, at scales of consideration ranging from individual facets of the urban environment, through streets and city blocks to neighbourhoods. Second, it explores the literature on the urban temperature field. The state of knowledge about urban heat islands around 1980 is described and work since then is assessed in terms of similarities to and contrasts with that situation. Finally, the main advances are summarized and recommendations for urban climate work in the future are made. Copyright 2003 Royal Meteorological Society.
    B枚hm R.,P. D. Jones,J. Hiebl,D. Frank,M. Brunetti, and M. Maugeri, 2010: The early instrumental warm-bias: A solution for long Central Europe an temperature series 1760-2007. Climatic Change, 101, 41-67.10.1007/s10584-009-9649-49e537410-788a-4101-95cd-59ec84a39cfbslarticleid_2001439cb09477dc7923b78c482fad7e3ee384http%3A%2F%2Fwww.springerlink.com%2Fcontent%2Fg111046235jnv572%2Frefpaperuri:(a03f19fe4167e6a8dcefe984bc726b6a)http://www.springerlink.com/content/g111046235jnv572/Instrumental temperature recording in the Greater Alpine Region (GAR) began in the year 1760. Prior to the 1850–1870 period, after which screens of different types protected the instruments, thermometers were insufficiently sheltered from direct sunlight so were normally placed on north-facing walls or windows. It is likely that temperatures recorded in the summer half of the year were biased warm and those in the winter half biased cold, with the summer effect dominating. Because the changeover to screens often occurred at similar times, often coincident with the formation of National Meteorological Services (NMSs) in the GAR, it has been difficult to determine the scale of the problem, as all neighbour sites were likely to be similarly affected. This paper uses simultaneous measurements taken for eight recent years at the old and modern site at Kremsmünster, Austria to assess the issue. The temperature differences between the two locations (screened and unscreened) have caused a change in the diurnal cycle, which depends on the time of year. Starting from this specific empirical evidence from the only still existing and active early instrumental measuring site in the region, we developed three correction models for orientations NW through N to NE. Using the orientation angle of the buildings derived from metadata in the station histories of the other early instrumental sites in the region (sites across the GAR in the range from NE to NW) different adjustments to the diurnal cycle are developed for each location. The effect on the 32 sites across the GAR varies due to different formulae being used by NMSs to calculate monthly means from the two or more observations made at each site each day. These formulae also vary with time, so considerable amounts of additional metadata have had to be collected to apply the adjustments across the whole network. Overall, the results indicate that summer (April to September) average temperatures are cooled by about 0.4°C before 1850, with winters (October to March) staying much the same. The effects on monthly temperature averages are largest in June (a cooling from 0.21° to 0.93°C, depending on location) to a slight warming (up to 0.3°C) at some sites in February. In addition to revising the temperature evolution during the past centuries, the results have important implications for the calibration of proxy climatic data in the region (such as tree ring indices and documentary data such as grape harvest dates). A difference series across the 32 sites in the GAR indicates that summers since 1760 have warmed by about 1°C less than winters.
    Bojinski S.,M. Verstraete,T. C. Peterson,C. Richter,A. Simmons, and M. Zemp, 2014: The concept of essential climate variables in support of climate research, applications, and policy. Bull. Amer. Meteor. Soc., 95, 1431-1443.10.1175/BAMS-D-13-00047.16bbc4b83a8bf904c556eea30675e8419http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F271271716_The_concept_of_Essential_Climate_Variables_in_support_of_climate_research_applications_and_policyhttp://www.researchgate.net/publication/271271716_The_concept_of_Essential_Climate_Variables_in_support_of_climate_research_applications_and_policyClimate research, monitoring, prediction, and related services rely on accurate observations of the atmosphere, land, and ocean, adequately sampled globally and over sufficiently long time periods. The Global Climate Observing System, set up under the auspices of United Nations organizations and the International Council for Science to help ensure the availability of systematic observations of climate, developed the concept of essential climate variables (ECVs). ECV data records are intended to provide reliable, traceable, observation-based evidence for a range of applications, including monitoring, mitigating, adapting to, and attributing climate changes, as well as the empirical basis required to understand past, current, and possible future climate variability. The ECV concept has been broadly adopted worldwide as the guiding basis for observing climate, including by the United Nations Framework Convention on Climate Change (UNFCCC), WMO, and space agencies operating Earth observation satellites. This paper describes the rationale for these ECVs and their current selection, based on the principles of feasibility, relevance, and cost effectiveness. It also provides a view of how the ECV concept could evolve as a guide for rational and evidence-based monitoring of climate and environment. Selected examples are discussed to highlight the benefits, limitations, and future evolution of this approach. The article is intended to assist program managers to set priorities for climate observation, dataset generation and related research: for instance, within the emerging Global Framework for Climate Services (GFCS). It also helps the observation community and individual researchers to contribute to systematic climate observation, by promoting understanding of ECV choices and the opportunities to influence their evolution.
    Brohan P.,J. J. Kennedy,I. Harris,S. F. B. Tett, and P. D. Jones, 2006: Uncertainty estimates in regional and global observed temperature changes: A new data set from 1850. J. Geophys. Res., 111, D12106, doi: 10.1029/2005JD006548.10.1029/2005JD0065487af8b0918e2d7a885a9114718b2a32d8http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2005JD006548%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2005JD006548/abstractABSTRACT The historical surface temperature data set HadCRUT provides a record of surface temperature trends and variability since 1850. A new version of this data set, HadCRUT3, has been produced, benefiting from recent improvements to the sea surface temperature data set which forms its marine component, and from improvements to the station records which provide the land data. A comprehensive set of uncertainty estimates has been derived to accompany the data: Estimates of measurement and sampling error, temperature bias effects, and the effect of limited observational coverage on large-scale averages have all been made. Since the mid twentieth century the uncertainties in global and hemispheric mean temperatures are small, and the temperature increase greatly exceeds its uncertainty. In earlier periods the uncertainties are larger, but the temperature increase over the twentieth century is still significantly larger than its uncertainty.
    Brunet M.,Coauthors, 2011: The minimization of the screen bias from ancient Western Mediterranean air temperature records: an exploratory statistical analysis. Inter.J. Climatol., 31, 1879-1895, doi: 10.1002/joc. 2192.
    Callendar G. S.,1938: The artificial production of carbon dioxide and its influence on temperature. Quart. J. Roy. Meteor. Soc., 64, 223-240, doi: 10.1002/qj.49706427503.10.1002/qj.49706427503f0137b51-05a3-4f9a-8808-d00be556087d8ef69664476d96de1f96321994dd0a71http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49706427503%2Fabstractrefpaperuri:(941b34c54bc01a23d6d96287516b6c21)http://onlinelibrary.wiley.com/doi/10.1002/qj.49706427503/abstractAbstract By fuel combustion man has added about 150,000 million tons of carbon dioxide to the air during the past half century. The author estimates from the best available data that approximately three quarters of this has remained in the atmosphere. The radiation absorption coefficients of carbon dioxide and water vapour are used to show the effect of carbon dioxide on “sky radiation.” From this the increase in mean temperature, due to the artificial production of carbon dioxide, is estimated to be at the rate of 0.003ºC. per year at the present time. The temperature observations a t zoo meteorological stations are used to show that world temperatures have actually increased at an average rate of 0.005ºC. per year during the past half century.
    Callendar G. S.,1961: Temperature fluctuations and trends over the earth. Quart. J. Roy. Meteor. Soc., 87, 1-12, doi: 10.1002/ qj.49708737102.10.1002/qj.497087373164c376810799fcea6a453c6c980d83aebhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49708737316%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1002/qj.49708737316/pdfABSTRACT The annual temperature deviations at over 400 meteorological stations are combined on a regional basis to give the integrated fluctuations over large areas and zones. These are shown in graphical form, and it is concluded that a solar or atmospheric dust hypothesis is necessary to explain the world-wide fluctuations of a few years duration. An important change in the relationships of the zonal fluctuations has occurred since 1920. The overall temperature trends found from the data are considered in relation to the homogeneity of recording, and also to the evidence of glacial recession in different zones. It is concluded that the rising trend, shown by the instruments during recent decades, is significant from the Arctic to about 45ºS lat., but quite small in most regions below 35ºN. and not yet apparent in some. It is thought that the regional and zonal distribution of recent climatic trends is incompatible with the hypothesis of increased solar heating as the cause. On the other hand, the major features of this distribution are not incompatible with the hypothesis of increased carbon dioxide radiation, if the rate of atmospheric mixing between the hemispheres is a matter of decades rather than years.
    Compo, G. P.,Coauthors, 2011: The twentieth century reanalysis project. Quart. J. Roy. Meteor. Soc., 137, 1-28, doi: 10.1002/qj.776.10.1002/qj.77670a19137-8c44-4632-86bd-2a4be97beb2afc49dba74377f61b8075326db5028f15http%3A%2F%2Fams.confex.com%2Fams%2F89annual%2Ftechprogram%2Fpaper_143366.htmhttp://ams.confex.com/ams/89annual/techprogram/paper_143366.htmThe Twentieth Century Reanalysis (20CR) project is an international effort to produce a comprehensive global atmospheric circulation dataset spanning the twentieth century, assimilating only surface pressure reports and using observed monthly sea-surface temperature and sea-ice distributions as boundary conditions. It is chiefly motivated by a need to provide an observational dataset with quantified uncertainties for validations of climate model simulations of the twentieth century on all time-scales, with emphasis on the statistics of daily weather. It uses an Ensemble Kalman Filter data assimilation method with background ‘first guess’ fields supplied by an ensemble of forecasts from a global numerical weather prediction model. This directly yields a global analysis every 6 hours as the most likely state of the atmosphere, and also an uncertainty estimate of that analysis.The 20CR dataset provides the first estimates of global tropospheric variability, and of the dataset's time-varying quality, from 1871 to the present at 6-hourly temporal and 2° spatial resolutions. Intercomparisons with independent radiosonde data indicate that the reanalyses are generally of high quality. The quality in the extratropical Northern Hemisphere throughout the century is similar to that of current three-day operational NWP forecasts. Intercomparisons over the second half-century of these surface-based reanalyses with other reanalyses that also make use of upper-air and satellite data are equally encouraging.It is anticipated that the 20CR dataset will be a valuable resource to the climate research community for both model validations and diagnostic studies. Some surprising results are already evident. For instance, the long-term trends of indices representing the North Atlantic Oscillation, the tropical Pacific Walker Circulation, and the Pacific–North American pattern are weak or non-existent over the full period of record. The long-term trends of zonally averaged precipitation minus evaporation also differ in character from those in climate model simulations of the twentieth century. Copyright 08 2011 Royal Meteorological Society and Crown Copyright.
    Compo G. P.,P. D. Sardesmukh,J. S. Whitaker,P. Brohan,P. D. Jones, and C. McColl, 2013: Independent confirmation of global land warming without the use of station. Geophys. Res. Lett., 40, 3170-3174, doi: 10.1002/grl.50425.
    Conrad V.,L. W. Pollak,1962: Methods in Climatology . Harvard University Press, 459 pp.
    Cowtan K.,R. G. Way, 2014: Coverage bias in the hadcrut4 temperature series and its impact on recent temperature trends. Quart. J. Roy. Meteor. Soc., 140, 1935-1944, doi: 10.1002/qj.2297.10.1002/qj.22975599281a-8e4d-4e2d-8771-1b5e3304cfc173f9abe155b0459562069164b14f2fe1http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.2297%2Ffullrefpaperuri:(6cef113a1ee34f3504ed57fd5bec2b64)http://onlinelibrary.wiley.com/doi/10.1002/qj.2297/fullAbstract Incomplete global coverage is a potential source of bias in global temperature reconstructions if the unsampled regions are not uniformly distributed over the planet's surface. The widely used Hadley Centre-揅limatic Reseach Unit Version 4 (HadCRUT4) dataset covers on average about 84% of the globe over recent decades, with the unsampled regions being concentrated at the poles and over Africa. Three existing reconstructions with near-global coverage are examined, each suggesting that HadCRUT4 is subject to bias due to its treatment of unobserved regions. Two alternative approaches for reconstructing global temperatures are explored, one based on an optimal interpolation algorithm and the other a hybrid method incorporating additional information from the satellite temperature record. The methods are validated on the basis of their skill at reconstructing omitted sets of observations. Both methods provide results superior to excluding the unsampled regions, with the hybrid method showing particular skill around the regions where no observations are available. Temperature trends are compared for the hybrid global temperature reconstruction and the raw HadCRUT4 data. The widely quoted trend since 1997 in the hybrid global reconstruction is two and a half times greater than the corresponding trend in the coverage-biased HadCRUT4 data. Coverage bias causes a cool bias in recent temperatures relative to the late 1990s, which increases from around 1998 to the present. Trends starting in 1997 or 1998 are particularly biased with respect to the global trend. The issue is exacerbated by the strong El Nino event of 1997-1998, which also tends to suppress trends starting during those years.
    Dee, D. P.,Coauthors, 2011: The ERA-Interim reanalysis: configuration and performance of the data assimilation system.Quart. J. Roy. Meteor. Soc., 137, 553-597, doi: 10.1002/ qj.828.10.1002/qj.828b8698c40-b145-4364-9b39-4e603f942b9f5e49541e9e977f77d4b4487298c60f84http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.828%2Fpdfrefpaperuri:(d4649bb38c91f047e85ec096d8587b99)http://onlinelibrary.wiley.com/doi/10.1002/qj.828/pdfABSTRACT ERA-Interim is the latest global atmospheric reanalysis produced by the European Centre for Medium-Range Weather Forecasts (ECMWF). The ERA-Interim project was conducted in part to prepare for a new atmospheric reanalysis to replace ERA-40, which will extend back to the early part of the twentieth century. This article describes the forecast model, data assimilation method, and input datasets used to produce ERA-Interim, and discusses the performance of the system. Special emphasis is placed on various difficulties encountered in the production of ERA-40, including the representation of the hydrological cycle, the quality of the stratospheric circulation, and the consistency in time of the reanalysed fields. We provide evidence for substantial improvements in each of these aspects. We also identify areas where further work is needed and describe opportunities and objectives for future reanalysis projects at ECMWF. Copyright 2011 Royal Meteorological Society
    Farmer G.,T. M. L. Wigley,P. D. Jones, and M. Salmon, 1989: Documenting and explaining recent global-mean temperature changes. Final Report to the Natural Environment Research Council, Contract No. GR3/6565, East Anglia University, Norwich,UK.52854fbfac59ea2a2207d9f9c0a1be8chttp%3A%2F%2Fueaeprints.uea.ac.uk%2F34251%2Fhttp://ueaeprints.uea.ac.uk/34251/
    Folland C. K.,2005: Assessing bias corrections in historical sea surface temperature using a climate model. Inter. J. Climatol., 25, 895-911, doi: 10.1002/joc.1171.10.1002/joc.1171ca9a3c2c5b8eb3798d9460d23470ed36http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.1171%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/joc.1171/fullAbstract Analyses of simulations of variations in global and large-regional land surface air temperature (LSAT) for 1872–1998 using the HadAM3 atmospheric general circulation model are reported. The analyses are designed to test the accuracy of bias corrections to sea-surface temperature (SST) used in the Hadley Centre's global sea ice and SST (GISST3.1) data set, the more recent Hadley Centre sea ice and SST (HadISST) data set, and in the underlying Met Office historical SST (MOHSST and HadSST1) data sets. The tests are important because SST corrections considerably affect estimates of the magnitude of global warming since the late 19th century. Two ensembles of simulations were created using GISST3.1 as the lower boundary condition. The first ensemble, of six integrations, was forced using GISST with bias-corrections applied from 1871 until 1941, and was continued with no bias corrections to 1998. A second ensemble of four integrations, for 1871 to 1941, was forced with uncorrected GISST data. Simulations with uncorrected GISST show a substantial and often highly significant cold bias in simulated global and large-regional annual mean LSAT changes before 1942 relative to a 1946–65 reference period. By contrast, corrected SST data led to simulations of LSAT changes that are generally insignificantly different from those of observed LSAT in most regions before 1942. Tests on extratropical hemispheric scales generally validate the seasonal variation of the bias corrections, though less clearly before 1890 in some seasons. Issues about the quality of the LSAT data are raised by the results in a couple of regions. Over Australia, the model may have reconstructed LSAT changes using bias-corrected GISST with greater accuracy than the observations before about 1910. Crown Copyright 2005. Reproduced with the permission of Her Majesty's Stationery Office. Published by John Wiley & Sons, Ltd.
    Folland C. K.,D. E. Parker, 1995: Correction of instrumental biases in historical sea surface temperature data. Quart. J.Roy. Meteor. Soc., 121, 319-367.10.1002/qj.49712152206bde503c087658692efe8b2da7accb67chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49712152206%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/qj.49712152206/abstractAbstract We describe a physically based empirical technique for correcting historical sea surface temperature measurements for time-varying biases. The corrections are based on models of heat and moisture transfers from uninsulated (canvas) and partially insulated (wooden) sea temperature buckets exposed on deck. One of the canvas bucket models is tested using measurements on board ship and published wind-tunnel measurements. The method gives geographically and seasonally varying bias corrections through the period 1856 to 1941. The corrections are fairly insensitive to uncertainties such as the size of the bucket or the details of its exposure on deck. A discussion of the history of sea surface temperature observations provides a background to the procedure. The resulting globally and seasonally averaged sea surface temperature corrections increase from 0.11 degC in 1856 to 0.42 degC by 1940. The corrections are compatible with recent measurements made at sea of the errors of canvas buckets. Global and hemispheric time series of corrected sea surface temperature and night marine air temperature data show good agreement: more detailed verifications of the corrections will be reported elsewhere.
    Foster G.,S. Rahmstorf, 2011: Global temperature evolution 1979-2010. Environ. Res. Lett., 6, 044022, doi: 10.1088/1748-9326/6/4/044022.10.1088/1748-9326/6/4/044022e67d5e28-c484-45ec-858d-dfcca4712ce1bb3b7b47910d95cf36f6f8cc29c2d40ehttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F254496419_Global_temperature_evolution_19792010%3Fev%3Dauth_pubrefpaperuri:(62cc21d0ae3f99b598aed2cccd520e8c)http://www.researchgate.net/publication/254496419_Global_temperature_evolution_19792010?ev=auth_pub. When the data are adjusted to remove the estimated impact of known factors on short-term temperature variations (El Nino/southern oscillation, volcanic aerosols and solar variability), the global warming signal becomes even more evident as noise is reduced. Lower-troposphere temperature responds more strongly to El Nino/southern oscillation and to volcanic forcing than surface temperature data. The adjusted data show warming at very similar rates to the unadjusted data, with smaller probable errors, and the warming rate is steady over the whole time interval. In all adjusted series, the two hottest years are 2009 and 2010.
    Hansen J.,R. Ruedy,J. Glascoe, and M. Sato, 1999: GISS analysis of surface temperature change. J. Geophys. Res., 104, 30 997-31 022, doi: 10.1029/1999JD900835.10.1029/1999JD900835616d2326-7ed6-4475-96a7-69eb66bc2291739f664db1de69c4fbde35a08320f21ehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F1999JD900835%2Ffullrefpaperuri:(b5c2460c0ce46f58754a152c32bfc827)http://onlinelibrary.wiley.com/doi/10.1029/1999JD900835/fullWe describe the current GISS analysis of surface temperature change for the period 1880–1999 based primarily on meteorological station measurements. The global surface temperature in 1998 was the warmest in the period of instrumental data. The rate of temperature change was higher in the past 25 years than at any previous time in the period of instrumental data. The warmth of 1998 was too large and pervasive to be fully accounted for by the recent El Nino. Despite cooling in the first half of 1999, we suggest that the mean global temperature, averaged over 2–3 years, has moved to a higher level, analogous to the increase that occurred in the late 1970s. Warming in the United States over the past 50 years has been smaller than in most of the world, and over that period there was a slight cooling trend in the eastern United States and the neighboring Atlantic Ocean. The spatial and temporal patterns of the temperature change suggest that more than one mechanism was involved in this regional cooling. The cooling trend in the United States, which began after the 1930s and is associated with ocean temperature change patterns, began to reverse after 1979. We suggest that further warming in the United States to a level rivaling the 1930s is likely in the next decade, but reliable prediction requires better understanding of decadal oscillations of ocean temperature.
    Hansen J.,M. Sato,R. Ruedy,K. Lo,D. W. Lea, and M. Medina-Elizade,2006: Global temperature change. Proceedings of the National Academy of Sciences of the United States of America, 103, 14 288-14 293.10.1073/pnas.0606291103170010185a429274-61cc-4df4-8350-2eeb9941a92f7c3357b99d8633a2fdad2fc2226c376dhttp%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM17001018refpaperuri:(9cd13cac9d240432aab6f315ba3cdb17)http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM17001018Global surface temperature has increased approximately 0.2 degrees C per decade in the past 30 years, similar to the warming rate predicted in the 1980s in initial global climate model simulations with transient greenhouse gas changes. Warming is larger in the Western Equatorial Pacific than in the Eastern Equatorial Pacific over the past century, and we suggest that the increased West-East temperature gradient may have increased the likelihood of strong El Ninos, such as those of 1983 and 1998. Comparison of measured sea surface temperatures in the Western Pacific with paleoclimate data suggests that this critical ocean region, and probably the planet as a whole, is approximately as warm now as at the Holocene maximum and within approximately 1 degrees C of the maximum temperature of the past million years. We conclude that global warming of more than approximately 1 degrees C, relative to 2000, will constitute "dangerous" climate change as judged from likely effects on sea level and extermination of species.
    Hansen J.,R. Ruedy,M. Sato, and K. Lo, 2010: Global surface temperature change. Rev. Geophys., 48,RG4004, doi: 10.1029/2010RG000345.10.1029/2010RG000345067def09-5355-4e00-b4a8-812e0c96c68f50b55a6514a75fb0273bfc4d9ef31adehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010RG000345%2Fpdfrefpaperuri:(f9fb10bbf865e1daf2c73f0d8e02dd4b)http://onlinelibrary.wiley.com/doi/10.1029/2010RG000345/pdfABSTRACT We update the Goddard Institute for Space Studies (GISS) analysis of global surface temperature change, compare alternative analyses, and address questions about perception and reality of global warming. Satellite-observed nightlights are used to identify measurement stations located in extreme darkness and adjust temperature trends of urban and peri-urban stations for non-climatic factors, verifying that urban effects on analyzed global change are small. Because the GISS analysis combines available sea surface temperature records with meteorological station measurements, we test alternative choices for the ocean data, showing that global temperature change is sensitive to estimated temperature change in polar regions where observations are limited. We use simple 12-month (and n×12) running means to improve the information content in our temperature graphs. Contrary to a popular misconception, the rate of warming has not declined. Global temperature is rising as fast in the past decade as in the prior two decades, despite year-to-year fluctuations associated with the El Nino-La Nina cycle of tropical ocean temperature. Record high global 12-month running-mean temperature for the period with instrumental data was reached in 2010.
    Hartmann, D. L.,Coauthors, 2013: Observations: Atmosphere and surface. Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, T. F. Stocker et al., Eds. Cambridge University Press.45d2d09c1905a19471963286594d9596http%3A%2F%2Feprints.soton.ac.uk%2F363409%2Fhttp://eprints.soton.ac.uk/363409/It is virtually certain that atmospheric burdens of long-lived greenhouse gases controlled by the Kyoto 6 Protocol increased from 2005 to 2011. Annual increases in global mean CO2 and N2O mole fractions were at 7 rates comparable to those
    Hawkins E.,P. D. Jones, 2013: On increasing global temperatures: 75 years after Callendar. Quart. J. Roy. Meteor. Soc., 139, 1961-1963, doi: 10.1002/qj.2178.10.1002/qj.2178a65e329f093683d6504b4950795c3ff5http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.2178%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1002/qj.2178/citedbyIn 1938, Guy Stewart Callendar was the first to demonstrate that the Earth's land surface was warming. Callendar also suggested that the production of carbon dioxide by the combustion of fossil fuels was responsible for much of this modern change in climate. This short note marks the 75th anniversary of Callendar's landmark study and demonstrates that his global land temperature estimates agree remarkably well with more recent analyses.
    Hersbach H.,C. Peubey,A. Simmons,P. Berrisford,P. Poli, and D. Dee, 2015: ERA-20CM: A twentieth-century atmospheric model ensemble. Quart. J. Roy. Meteor. Soc., 141, 2350-2375, doi: 10.1002/qj.2528.10.1002/qj.2528a24465adfa39e6e0cd593229f1c8ed03http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.2528%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1002/qj.2528/citedbyABSTRACT This paper describes an ensemble of ten atmospheric model integrations for the years 1899 to 2010, performed at the European Centre for Medium-Range Weather Forecasts (ECMWF). Horizontal spectral resolution is T159 (about 125 km), using 91 levels in the vertical from the surface up to 1 Pa, and a time step of one hour. This ensemble, denoted by ERA-20CM, formed the first step toward a 20th century reanalysis within ERA-CLIM, a three-year European funded project involving nine partners.Sea-surface temperature and sea-ice cover are prescribed by an ensemble of realizations (HadISST2), as recently produced by the Met Office Hadley Centre within ERA-CLIM. Variation in these realizations reflect uncertainties in the available observational sources on which this product is based. Forcing terms in the model radiation scheme follow CMIP5 recommendations. Any effect of their uncertainty is neglected. These terms include solar forcing, greenhouse gases, ozone and aerosols. Both the ocean-surface and radiative forcing incorporate a proper long-term evolution of climate trends in the 20th century, and the occurrence of major events, such as the El Niño-Southern Oscillations and volcanic eruptions.No atmospheric observations were assimilated. For this reason ERA-20CM is not able to reproduce actual synoptic situations. The ensemble is, however, able to provide a statistical estimate of the climate. Overall, the temperature rise over land is in fair agreement with the CRUTEM4 observational product. Over the last two decades the warming over land exceeds the warming over sea, which is consistent with models participating in the CMIP5 project, as well with the ECMWF ERA-Interim reanalysis. Some aspects of warming and of the hydrological cycle are discerned, and the model response to volcanic eruptions is qualitatively correct.The data from ERA-20CM are freely available, embracing monthly-mean fields for many atmospheric and ocean-wave quantities, and synoptic fields for a small, essential subset.
    Huang B.Y.,Coauthors, 2015: Extended reconstructed Sea surface temperature Version 4 (ERSST.v4). Part I: upgrades and intercomparisons. J.Climate, 28, 911-930.10.1175/JCLI-D-14-00006.138f80fcf7c4c7442e8596106a00e6b20http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F272403298_Extended_Reconstructed_Sea_Surface_Temperature_Version_4_%28ERSST.v4%29._Part_I_Upgrades_and_Intercomparisonshttp://www.researchgate.net/publication/272403298_Extended_Reconstructed_Sea_Surface_Temperature_Version_4_(ERSST.v4)._Part_I_Upgrades_and_IntercomparisonsABSTRACT The monthly Extended Reconstructed Sea Surface Temperature (ERSST) dataset, available on global 2 degrees x 2 degrees grids, has been revised herein to version 4 (v4) from v3b. Major revisions include updated and substantially more complete input data from the International Comprehensive Ocean Atmosphere Data Set (ICOADS) release 2.5; revised empirical orthogonal teleconnections (EOTs) and EOT acceptance criterion; updated sea surface temperature (SST) quality control procedures; revised SST anomaly (SSTA) evaluation methods; updated bias adjustments of ship SSTs using the Hadley Centre Nighttime Marine Air Temperature dataset version 2 (HadNMAT2); and buoy SST bias adjustment not previously made in v3b. Tests show that the impacts of the revisions to ship SST bias adjustment in ERSST.v4 are dominant among all revisions and updates. The effect is to make SST 0.1 degrees-0.2 degrees C cooler north of 30 degrees S but 0.1 degrees-0.2 degrees C warmer south of 30 degrees S in ERSST.v4 than in ERSST.v3b before 1940. In comparison with the Met Office SST product [the Hadley Centre Sea Surface Temperature dataset, version 3 (HadSST3)], the ship SST bias adjustment in ERSST.v4 is 0.1 degrees-0.2 degrees C cooler in the tropics but 0.1 degrees-0.2 degrees C warmer in the midlatitude oceans both before 1940 and from 1945 to 1970. Comparisons highlight differences in long-term SST trends and SSTA variations at decadal time scales among ERSST.v4, ERSST.v3b, HadSST3, and Centennial Observation-Based Estimates of SST version 2 (COBE-SST2), which is largely associated with the difference of bias adjustments in these SST products. The tests also show that, when compared with v3b, SSTAs in ERSST.v4 can substantially better represent the El Nino/La Nina behavior when observations are sparse before 1940. Comparisons indicate that SSTs in ERSST.v4 are as close to satellite-based observations as other similar SST analyses.
    Ishii M.,A. Shouji,S. Sugimoto, and T. Matsumoto, 2005: Objective analyses of Sea-surface temperature and marine meteorological variables for the 20th century using ICOADS and the Kobe collection. Inter. J.Climatol., 25, 865-879.10.1002/joc.11693d60cc42-6778-4fb8-9700-e3d7903911024646cedb71328599c9da066082123167http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.1169%2Ffullrefpaperuri:(9346fce3aa27305acc6886acb15f9b78)http://onlinelibrary.wiley.com/doi/10.1002/joc.1169/fullAbstract Data for the 20th century from the International Comprehensive Ocean and Atmosphere Data Set and the Kobe Collection have been used as input data for global objective analyses of sea-surface temperatures (SSTs) and other marine meteorological variables. This study seeks a better understanding of the historical marine meteorological data and an evaluation of the quality of the data in the Kobe Collection. Objective analyses yield gridded data that are less noisy than observed data, which facilitates handling of historical data. The observed data determine the quality of the objective analyses, and quality control specified for historical data is incorporated into the objective analysis to reduce artificial errors. The objective analyses are based on optimum interpolation and reconstruction with empirical orthogonal functions. The final database produced in this study not only contains analysed values, but also analysis errors and data distributions at each time step of the objective analyses. The analysis database contains ample information on historical observations, as well as signals of marine climate variations during the century. Time series of global mean marine temperatures and cloud cover include trends linked to global warming, and local peaks appear commonly in all the time series around the 1940s. Sea-level pressure and sea-surface wind fields show significant linear trends at high latitudes and over the North Pacific Ocean respectively. These trends seem to be artificial. An SST analysis used widely in climatological studies was verified against HadISST from the Hadley Centre and an SST analysis derived from satellite and in situ observations. El Niño and southern oscillation indices for the century are successfully reproduced, even though observations in the tropics are much rarer before 1950 than after 1950. Copyright 2005 Royal Meteorological Society
    Jansen E.,Coauthors, 2007: Palaeoclimate. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, S. Solomon et al., Eds. Cambridge University Press, 433-497.b563404398fbb06b53855528b8e05782http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F272791501_Intergovernmental_Panel_on_Climate_ChangeContribution_of_Working_Group_I_to_the_Fourth_Assessment_Report_of_the_Intergovernmental_Panel_on_Climate_Change._Cambridge_University_Press_Cambridge_and_New_York_http://www.researchgate.net/publication/272791501_Intergovernmental_Panel_on_Climate_ChangeContribution_of_Working_Group_I_to_the_Fourth_Assessment_Report_of_the_Intergovernmental_Panel_on_Climate_Change._Cambridge_University_Press_Cambridge_and_New_York_
    Jones P. D.,1994: Hemispheric surface air temperature variations: a reanalysis and an update to 1993. J.Climate, 7, 1794-1802.10.1175/1520-0442(1994)007<1794:HSATVA>2.0.CO;284ac8eeb2fb7e4d30e3ff87bb48979bchttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F236401573_Hemispheric_surface_air_temperature_variations_A_reanalysis_and_an_update_to_1993http://www.researchgate.net/publication/236401573_Hemispheric_surface_air_temperature_variations_A_reanalysis_and_an_update_to_1993Abstract Land-based compilations of gridded monthly surface air temperature anomalies, averaged into hemispheric values for the last 140 years, have been available for climatological analyses for the last 10 years or so. The analysis techniques used in their construction, particularly the need for a common reference period, mean that it is difficult to include, retrospectively, any of the new temperature datasets now available for some countries. So, despite data availability improvements in some areas, the number of stations used has fallen since 1970, both in the hemispheric averages and in their constituent grid-box datasets. The present study is a reanalysis of both the existing and the newly available temperature datasets to produce a grid-box dataset of 5º5º temperature anomalies. The reanalysis not only uses over 1000 more stations (2961 in total), principally covering the period from the 1920s to about 1990, but also arrests the decline of stations incorporated in real time for the latest years. Two hundred and fifty-two more stations are used in this analysis for the 1991-1993 period, compared with earlier analyses, The purpose of the reanalysis, however, is not just to calculate hemispheric averages. The improvements in station numbers used mean that the grid-box dataset should better estimate time series for small subcontinental scales. Despite the dramatic improvements in the numbers of stations used, the results change little from earlier analyses for the Northern Hemisphere average, indicating the robustness of the earlier time series. Similar results could have been achieved with as few as 109 stations. Over the Southern Hemisphere, comparisons of the results indicate larger (but still relatively small) differences with earlier analyses, particularly over continental-scale regions.
    Jones P. D.,D. H. Lister, 2009: The urban heat island in central London and urban-related warming trends in central London since 1900. Weather, 64, 323-327.10.1002/wea.4324bf34ce54d1c424fa8c5130b38556d4ahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fwea.432%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/wea.432/fullWhile flowing through porous medium, groundwater flow dissolves minerals thereby increasing medium porosity and Ultimately permeability. Reactive fluid flows preferentially into highly permeable zones, which are therefore dissolved most rapidly, producing a further preferential permeability enhancement. Accordingly, slight non-uniformities present in porous medium can be amplified and lead to fingering reaction fronts. The objective of this Study is to investigate dissolution-induced porosity changes on reaction front morphology In homogeneous porous medium With two non-uniformities. Four Controlling parameters, including upstream pressure gradient, reaction rate constant, non-uniformities spacing and non-uniformity strength ratio are comprehensively considered. By using a modified version of the numerical code, NSPCRT, to conduct a series of numerical simulations, front behavior diagrams are constructed to illustrate the morphologies of reaction fronts under various combinations of these four factors. Simulation results indicate that the two non-uniformities are inhibited into a planar front under low Upstream pressure gradient, merge into a single-fingering front under intermediate upstream pressure gradient, or grow into a double-fingers front under high upstream pressure gradient. Moreover, the two non-uniformities tend to develop into a double-fingering front as the non-uniformity strength ratio increases from 0.2 to 1.0, and merge into a single-fingering front while the non-uniformity strength ratio increases From 1.0 to 1.8. When the reaction rate constant is small, the two non-uniformities merge into a single front. Reaction rate constant significantly affects front advancing velocity. The front advancing velocity decreases with the reaction rate constant. Based on these results, front behavior diagrams which define the morphologies of the reaction fronts for these four parameters are constructed. Moreover, non-uniformity strength ratio and reaction rate constant are identified as two important factors that govern the interaction of dissolution and Solute transport in groundwater systems.
    Jones P. D.,D. H. Lister, 2015: Antarctic near-surface air temperatures compared with ERA-Interim values since 1979. International Journal of Climatology, 35, 1354-1366, doi: 10.1002/joc.4061.10.1002/joc.4061a4fe6c0cf7367cfbb2c44a6fc428e02chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.4061%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/joc.4061/fullABSTRACT ERA-Interim reanalysis for near-surface air temperature agrees well with land stations in most regions with the principal exception of the Antarctic. Here we compare annual and monthly values from 40 manned and automatic weather stations (AWSs) with ERA-Interim between 1979 and 2013. In terms of absolute differences between ERA-Interim and the station observations, ERA-Interim is biased warm (by up to 5 °C) at the few inland stations, but biased cool at lower latitudes between 65°S and 78°S (by up to 6 °C) at some locations. These biases tend to reduce between the period 1979&ndash;1990 and 2002&ndash;2013 at many sites, but they increase at three sites on the Antarctic Peninsula. Comparisons of differences in variability show one or nine stations have standard deviation values for ERA-Interim 20% less or more than the station series. Time series agreement in terms of monthly correlations of anomalies is good (r ≥ 0.89) for most of the stations, but seven are below 0.80, with ten between the two thresholds. Finally, we produce an &lsquo;Antarctic&rsquo; average time series by simple averaging of the 40 stations and ERA-Interim time series, as well as calculating an average based on all land points across the Antarctic from ERA-Interim. The series based on all land points is more variable on the year-to-year timescale and trends for the overall period are reduced.
    Jones P. D.,T. M. L. Wigley, 2010: Estimation of global temperature trends: What important and what isn. Climatic Change, 100, 59-69.
    Jones P. D.,P. Y. Groisman,M. Coughlan,N. Plummer,W.-C. Wang, and T. R. Karl, 1990: Assessment of urbanization effects in time series of surface air temperature over land. Nature, 347, 169-172.10.1038/347169a05c54ca5f7fd508b7580eb304a22f2a15http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F19911952088.htmlhttp://www.cabdirect.org/abstracts/19911952088.htmlRecords of hemisphere average temperatures from land regions for the past 100 years provide crucial input to the debate over global warming. Despite careful use of the basic station data there have been suggestions that a proportion of the 0.5ºC warming seen on a century timescale may be related to urbanization influence. An extensive set of rural-station temperature data for three regions of t...
    Jones P. D.,T. J. Osborn, and K. R. Briffa, 1997: Estimating sampling errors in large-scale temperature averages. J.Climate, 10, 2548-2568.10.1175/1520-0442(1997)010<2548:ESEILS>2.0.CO;2c114f7cf73275e9d36e4eec566cb41f6http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F228765281_Estimating_sampling_errors_in_large-scale_temperature_averageshttp://www.researchgate.net/publication/228765281_Estimating_sampling_errors_in_large-scale_temperature_averagesAbstract A method is developed for estimating the uncertainty (standard error) of observed regional, hemispheric, and global-mean surface temperature series due to incomplete spatial sampling. Standard errors estimated at the grid-box level [SE 2 = S 2 (1 61 r00 )/(1 + ( n 61 1) r00 )] depend upon three parameters: the number of site records ( n ) within each box, the average interrecord correlation ( r00 ) between these sites, and the temporal variability ( S 2 ) of each grid-box temperature time series. For boxes without data ( n = 0), estimates are made using values of S 2 interpolated from neighboring grid boxes. Due to spatial correlation, large-scale standard errors in a regional-mean time series are not simply the average of the grid-box standard errors, but depend upon the effective number of independent sites ( N eff ) over the region. A number of assumptions must be made in estimating the various parameters, and these are tested with observational data and complementary results from multicentury control integrations of three coupled general circulation models (GCMs). The globally complete GCMs enable some assumptions to be tested in a situation where there are no missing data; comparison of parameters computed from the observed and model datasets are also useful for assessing the performance of GCMs. As most of the parameters are timescale dependent, the resulting errors are likewise timescale dependent and must be calculated for each timescale of interest. The length of the observed record enables uncertainties to be estimated on the interannual and interdecadal timescales, with the longer GCM runs providing inferences about longer timescales. For mean annual observed data on the interannual timescale, the 95% confidence interval for estimates of the global-mean surface temperature since 1951 is ±0.12°C. Prior to 1900, the confidence interval widens to ±0.18°C. Equivalent values on the decadal timescale are smaller: ±0.10°C (1951–95) and ±0.16°C (1851–1900).
    Jones P. D.,K. R. Briffa, and T. J. Osborn, 2003: Changes in the Northern hemisphere annual cycle: Implications for paleoclimatology? J.Geophys. Res., 108, 4588, doi: 10.1029/2003JD 003695.10.1029/2003JD003695da4decb5c21848b28be9ff5d19bf6be6http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2003JD003695%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2003JD003695/abstract[1] Paleoclimatologists generally consider past epochs on the basis of whether they were warmer or colder than today's climate. It is often not possible, however, to consider potential changes in the annual cycle because of limited seasonal emphases in many climate proxies. Using both long European instrumental records and longer European and Chinese documentary series, we show that winters have warmed relative to summers over the last 200 years compared to earlier in the millennium. Without more widespread data we do not know how general these changes are, but if they are applicable to other parts of the world, there are two principal implications for paleoclimatology. First, because high-frequency climate information obtained from some proxies is more indicative of “summer” conditions, it may give erroneous indications of the past if used to reconstruct mean temperatures over the whole calendar year. Second, climate model integrations of the last millennium would be expected to produce seasonal differences on century timescales, so this measure should be used as one of the important indicators of model performance.
    Jones P. D.,D. H. Lister, and Q. Li, 2008: Urbanization effects in large-scale temperature records, with an emphasis on China. J. Geophys. Res., 113, D16122, doi: 10.1029/2008JD009916.10.1029/2008JD0099164ab0627c-7d99-426a-9bd6-4a95653ea183a72ecdb5823d519a3c1c5c408bf824c1http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2008JD009916%2Fpdfrefpaperuri:(02c52d263142e1b9f1420fcf4f46fff8)http://onlinelibrary.wiley.com/doi/10.1029/2008JD009916/pdf[1] &nbsp;Global surface temperature trends, based on land and marine data, show warming of about 0.8&deg;C over the last 100 years. This rate of warming is sometimes questioned because of the existence of well-known Urban Heat Islands (UHIs). We show examples of the UHIs at London and Vienna, where city center sites are warmer than surrounding rural locations. Both of these UHIs however do not contribute to warming trends over the 20th century because the influences of the cities on surface temperatures have not changed over this time. In the main part of the paper, for China, we compare a new homogenized station data set with gridded temperature products and attempt to assess possible urban influences using sea surface temperature (SST) data sets for the area east of the Chinese mainland. We show that all the land-based data sets for China agree exceptionally well and that their residual warming compared to the SST series since 1951 is relatively small compared to the large-scale warming. Urban-related warming over China is shown to be about 0.1&deg;C decade &minus;1 over the period 1951&ndash;2004, with true climatic warming accounting for 0.81&deg;C over this period.
    Jones P. D.,D. H. Lister,T. J. Osborn,C. Harpham,M. Salmon, and C. P. Morice, 2012: Hemispheric and large-scale land-surface air temperature variations: An extensive revision and an update to 2010. J. Geophys. Res., 117, D05127, doi: 10.1029/2011JD017139.10.1029/2011JD017139fe02ce3f0880674dc8522d19bf329a3chttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F258662924_Hemispheric_and_large-scale_land-surface_air_temperature_variations_An_extensive_revision_and_an_update_to_2010http://www.researchgate.net/publication/258662924_Hemispheric_and_large-scale_land-surface_air_temperature_variations_An_extensive_revision_and_an_update_to_2010ABSTRACT
    Kalnay E.,Coauthors, 1996: The NCEP/NCAR 40-year reanalysis project. Bull. Amer. Meteor. Soc., 77, 437-471.23d674534321ec5c56bf181fd85f5561http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1536-1098-69-2-93-Kalnay1%26dbid%3D16%26doi%3D10.3959%252F1536-1098-69.2.93%26key%3D10.1175%252F1520-0477%281996%29077%3C0437%253ATNYRP%3E2.0.CO%253B2/s?wd=paperuri%3A%28fe1c070047a030c900beb40441caee5a%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1536-1098-69-2-93-Kalnay1%26dbid%3D16%26doi%3D10.3959%252F1536-1098-69.2.93%26key%3D10.1175%252F1520-0477%281996%29077%253C0437%253ATNYRP%253E2.0.CO%253B2&ie=utf-8
    Karl T. R.,C. N. Williams Jr., P. J. Young, and W. M. Wendland, 1986: A model to estimate the time of observation bias associated with monthly mean maximum, minimum and mean temperatures for the United States. J.Climate Appl. Meteor., 25, 145-160.10.1175/1520-0450(1986)0252.0.CO;27efed02f9be15377cb0ce1af5044d64bhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F234209511_A_Model_to_Estimate_the_Time_of_Observation_Bias_Associated_with_Monthly_Mean_Maximum_Minimum_and_Mean_Temperatures_for_the_United_Stateshttp://www.researchgate.net/publication/234209511_A_Model_to_Estimate_the_Time_of_Observation_Bias_Associated_with_Monthly_Mean_Maximum_Minimum_and_Mean_Temperatures_for_the_United_StatesAbstract Hourly data for 79 stations in the United States are used to develop an empirical model which can be used to estimate the time of observation bias associated with different observation schedules. The model is developed for both maximum and minimum monthly average temperature as well as monthly mean temperature. The model was tested on 28 independent stations, and the results were very good. Using seven years of hourly data the standard errors of estimate using the model were only moderately higher than the standard errors of estimate of the true time of observation bias. The physical characteristics of the model directly include a measure of mean monthly interdiurnal temperature differences, analemma information, and the effects of the daily temperature range due to solar forcing. A self-contained computer program has been developed which allows a user to estimate the time of observation bias anywhere in the contiguous United States without the costly exercise of accusing 24-hourly observations at first-order stations.
    Karl T. R.,R. W. Knight, and J. R. Christy, 1994: Global and hemispheric temperature trends: Uncertainties related to inadequate spatial sampling. J.Climate, 7, 1144-1163.10.1175/1520-0442(1994)0072.0.CO;2d3cca26daf7258942bfbca4097ae7d9bhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F236365736_Global_and_hemispheric_temperature_trends_Uncertainties_related_to_inadequate_spatial_samplinghttp://www.researchgate.net/publication/236365736_Global_and_hemispheric_temperature_trends_Uncertainties_related_to_inadequate_spatial_samplingAbstract Long-term (50 to 100 years) and short-term (10 to 30 years) global and hemispheric trends of temperature have an inherent unknown error due to incomplete and nonrandom spatial sampling. A number of experiments have been conducted to help quantify the potential magnitude of this error. The analysis includes the errors introduced into the climate record because of both incomplete global coverage and inadequate sampling within those areas presumed to have adequate observatory. In these experiments it is found that the uncertainty in calculating historical temperature trends is dependent upon the pattern of temperature change, the method of treating the effect of nonrandom spatial sampling, and the time and length over which the trend is calculated but is relatively insensitive to the random errors associated with estimating regional-scale (grid cell size) temperature anomalies. Results imply that the errors associated with century-scale trends of temperature are probably an order of magnitude smaller than the observed global warming of nearly 0.5ºC per 100 years since the late nineteenth century. The errors in estimates of decadal temperature trends are found to be large relative to century-scale trends and are unreliable during the nineteenth and early twentieth centuries. Even during the recent decade of the 1980s, the area-averaging techniques used in some analyses could be improved by addressing the over-sampling of Northern Hemisphere (especially over land) relative to the rest of the globe. Otherwise, significant positive biases are likely during the 1980s. These biases may have contributed to the reported differences between in situ surface and space-based temperatures during the 1980s. The rather encouraging results with respect to the magnitude of the spatial sampling errors associated with the calculation of long-term trends beginning in the nineteenth century cast a positive light on efforts aimed at extending the proxy and observed global temperature record further back in time, despite limited geographic coverage.
    Karl, T. R.,Coauthors, 2015: Possible artifacts of data biases in the recent global surface warming hiatus. Science, 348, 1469-1472.10.1126/science.aaa56325d986af30634bbe7f592f397edde3067http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F277779273_Possible_artifacts_of_data_biases_in_the_recent_global_surface_warming_hiatushttp://www.researchgate.net/publication/277779273_Possible_artifacts_of_data_biases_in_the_recent_global_surface_warming_hiatusMuch study has been devoted to the possible causes of an apparent decrease in the upward trend of global surface temperatures since 1998, a phenomenon that has been dubbed the global warming “hiatus.” Here, we present an updated global surface temperature analysis that reveals that global trends are higher than those reported by the Intergovernmental Panel on Climate Change, especially in recent decades, and that the central estimate for the rate of warming during the first 15 years of the 21st century is at least as great as the last half of the 20th century. These results do not support the notion of a “slowdown” in the increase of global surface temperature.
    Kennedy J. J.,2014: A review of uncertainty in in situ measurements and data sets of Sea surface temperature. Rev. Geophys., 52, 1-32, doi: 10.1002/2013RG000434.10.1002/2013RG0004345aa35066-3728-4818-ae7d-eb746397d2638322e5fba6b28077e5db17f28e9de1b0http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F2013RG000434%2Fabstractrefpaperuri:(4b0a6c283888ccdc67fa401679e2c9f3)http://onlinelibrary.wiley.com/doi/10.1002/2013RG000434/abstractArchives of in situ sea surface temperature (SST) measurements extend back more than 160 years. Quality of the measurements is variable, and the area of the oceans they sample is limited, especially early in the record and during the two world wars. Measurements of SST and the gridded data sets that are based on them are used in many applications so understanding and estimating the uncertainties are vital. The aim of this review is to give an overview of the various components that contribute to the overall uncertainty of SST measurements made in situ and of the data sets that are derived from them. In doing so, it also aims to identify current gaps in understanding. Uncertainties arise at the level of individual measurements with both systematic and random effects and, although these have been extensively studied, refinement of the error models continues. Recent improvements have been made in the understanding of the pervasive systematic errors that affect the assessment of long-term trends and variability. However, the adjustments applied to minimize these systematic errors are uncertain and these uncertainties are higher before the 1970s and particularly large in the period surrounding the Second World War owing to a lack of reliable metadata. The uncertainties associated with the choice of statistical methods used to create globally complete SST data sets have been explored using different analysis techniques, but they do not incorporate the latest understanding of measurement errors, and they want for a fair benchmark against which their skill can be objectively assessed. These problems can be addressed by the creation of new end-to-end SST analyses and by the recovery and digitization of data and metadata from ship log books and other contemporary literature.
    Kennedy J. J.,N. A. Rayner,R. O. Smith,D. E. Parker, and M. Saunby, 2011a: Reassessing biases and other uncertainties in Sea surface temperature observations measured in situ since 1850: 1. Measurement and sampling uncertainties. J. Geophys. Res., 116, doi: 10.1029/2010JD015218.10.1029/2010JD0152188dbbaec2ed5d93aab4c0c1b147b81b44http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010JD015218%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1029/2010JD015218/pdf[1] New estimates of measurement and sampling uncertainties of gridded in situ sea surface temperature anomalies are calculated for 1850 to 2006. The measurement uncertainties account for correlations between errors in observations made by the same ship or buoy due, for example, to miscalibration of the thermometer. Correlations between the errors increase the estimated uncertainties on grid box averages. In grid boxes where there are many observations from only a few ships or drifting buoys, this increase can be large. The correlations also increase uncertainties of regional, hemispheric, and global averages above and beyond the increase arising solely from the inflation of the grid box uncertainties. This is due to correlations in the errors between grid boxes visited by the same ship or drifting buoy. At times when reliable estimates can be made, the uncertainties in global average, Southern Hemisphere, and tropical sea surface temperature anomalies are between 2 and 3 times as large as when calculated assuming the errors are uncorrelated. Uncertainties of Northern Hemisphere averages are approximately double. A new estimate is also made of sampling uncertainties. They are largest in regions of high sea surface temperature variability such as the western boundary currents and along the northern boundary of the Southern Ocean. The sampling uncertainties are generally smaller in the tropics and in the ocean gyres.
    Kennedy J. J.,N. A. Rayner,R. O. Smith,D. E. Parker, and M. Saunby, 2011b: Reassessing biases and other uncertainties in Sea surface temperature observations measured in situ since 1850: 2. Biases and homogenization. J. Geophys. Res., 116, doi: 10.1029/2010JD015220.10.1029/2010JD0152204999c6e18caef979dba8569a5185c68ahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010JD015220%2Fsuppinfohttp://onlinelibrary.wiley.com/doi/10.1029/2010JD015220/suppinfoABSTRACT Errors in SST measurements are correlatedErrors in SST measurements have previously been underestimatedNew error estimates in SST data have been made which account for these
    Kent E. C.,J. J. Kennedy,D. I. Berry, and R. O. Smith, 2010: Effects of instrumentation changes on sea surface temperature measured in situ. Wiley Interdisciplinary Reviews: Climate Change, 1(5), 718-728, doi: 10.1002/wcc.55.10.1002/wcc.55bef7c2d839d532b93a91bcaf7582a826http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fwcc.55%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/wcc.55/abstractAbstract Measurements of sea surface temperature (SST) are an important climate record, complementing terrestrial air temperature observations, records of marine air temperature, ocean subsurface temperature, and ocean heat content. SST has been measured since the 18th century, although observations are sparse in the early period. Historically, marine observing systems relied on observations made by seafarers and necessary information on measurement methods is often not available. There are many historical descriptions of observing practice and instrumentation, some including quantification of biases between different methods. This documentation has been used, with the available observations, to develop models for the expected biases, which vary according to how the measurements were made, over time and with the environmental conditions. Adjustments have been developed for these biases and some gridded SST datasets adjust for these differences and provide uncertainty estimates, including uncertainties in the bias adjustments. The modern in situ SST-observing system continues to evolve and now includes many observations from moored and drifting buoys, which must be characterized relative to earlier observations to provide a consistent record of multi-decadal changes in SST. Copyright 2010 John Wiley & Sons, Ltd. For further resources related to this article, please visit the TODO: clickthrough URL WIREs website
    Kent E. C.,N. A. Rayner,D. I. Berry,M. Saunby,B. I. Moat,J. J. Kennedy, and D. E. Parker, 2013: Global analysis of night marine air temperature and its uncertainty since 1880: The HadNMAT2 data set. J. Geophys. Res., 118, 1281-1298, doi: 10.1002/jgrd.50152.10.1002/jgrd.50152e7832a469718ea9210343c26cd810147http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjgrd.50152%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/jgrd.50152/abstractABSTRACT An updated version of the Met Office Hadley Centre&rsquo;s monthly night marine air temperature dataset is presented. It is available on a 5藲 latitude-longitude grid from 1880 as anomalies relative to 1961-1990 calendar-monthly climatological average night marine air temperature (NMAT). Adjustments are made for changes in observation height; these depend on estimates of the stability of the near surface atmospheric boundary layer. In previous versions of the dataset, ad hoc adjustments were also made for three periods and regions where poor observational practice was prevalent. These adjustments are re-examined. Estimates of uncertainty are calculated for every grid box and result from: measurement errors; uncertainty in adjustments applied to the observations; uncertainty in the measurement height and under-sampling. The new dataset is a clear improvement over previous versions in terms of coverage because of the recent digitization of historical observations from ships' logbooks. However, the periods prior to about 1890 and around World War 2 remain particularly uncertain and sampling is still sparse in some regions in other periods. A further improvement is the availability of uncertainty estimates for every grid box and every month. Previous versions required adjustments that were dependent on contemporary measurements of sea surface temperature (SST); to avoid these, the new dataset starts in 1880 rather than 1856. Overall agreement with variations of SST is better for the updated dataset than for previous versions, maintaining existing estimates of global warming and increasing confidence in the global record of temperature variability and change.
    Köppen W.,1873: Über mehrjährige perioden der witterung, insbesondere über die 11-jährige periode der temperatur. Zeitschrift der Österreichischen Gesellschaft für Meteorologie, Bd VIII, 241-248, 257-267.
    Le Treut, H., R. Somerville, U. Cubasch, Y. Ding, C. Mauritzen, A. Mokssit, T. Peterson, and M. Prather, 2007: Historical overview of climate change. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, S. Solomon et al., Eds. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 93-127.10.1002/0471776289.ch1197948c8-47bf-46d1-b2e8-7aac4e02d8d018abc7b892cc0df17ec412f647083851http%3A%2F%2Fwww.citeulike.org%2Fgroup%2F13619%2Farticle%2F7673473refpaperuri:(767219252f43651afc25286784bafcba)http://www.citeulike.org/group/13619/article/7673473In this article, the author presents a historical overview of four decades of scholarship and changing public policy on family and informal caregiving for older adults. Families are changing at a dizzying pace. Changes in family composition, cultural diversity, geographic mobility, and societal norms, coupled with increasing numbers of older adults living with high levels of disability, are changing how caregiving for older adults is balanced among families, informal networks, and formal supports. Social policy, practice models, and empirical research have not kept pace with these changes. This country has yet to develop a comprehensive, integrated, long-term care system that views informal caregivers both as care partners and as service recipients in their own right. Moreover, recent policy changes designed to reduce government expenditures put families at risk for having to take on even greater care responsibilities. For better and for worse, communication and technologic interventions are gradually replacing some forms of direct human contact. It is still unclear whether older adults and their caregivers will become more isolated or whether new social contracts within families and between families and society will arise, rooted in the belief that even disabled older adults have a contribution to make, that they deserve to have their basic needs met, and that families should not be solely responsible for meeting those needs.
    Li Q.X, J. Y. Huang, Z. H. Jiang, L. M. Zhou, P. Chu,K. X. Hu, 2014: Detection of urbanization signals in extreme winter minimum temperature changes over Northern China. Climatic Change, 122, 595-608.10.1007/s10584-013-1013-z0195b31ecbf5ea2a33cbe6364ad6d79chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1007%2Fs10584-013-1013-zhttp://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1007/s10584-013-1013-zAlthough previous studies show that urbanization contributes to less than 10 of the long-term regional total warming trend of mean surface air temperature in northeast China (Li et al. 2010 ), the urban heat island (UHI) impact on extreme temperatures could be more significant. This paper examines the urbanization impact on extreme winter minimum temperatures from 33 stations in North China during the period of 1957-2010. We use the Generalized Extreme Value (GEV) distribution to analyze the distribution of extreme minimum temperatures and the long-term variations of the three distributional characteristics parameters. Results suggest that among the three distribution parameters, the position parameter is the most representative in terms of the long-term extreme minimum temperature change. A new classification method based on the intercommunity (factors analysis method) of the temperature change is developed to detect the urbanization effect on winter extreme minimum temperatures in different cities. During the period of rapid urbanization (after 1980), the magnitude of variations of the three distribution parameters for the urban station group is larger than that for the reference station group, indicating a higher chance of occurrence of warmer weather and a larger fluctuation of temperatures. Among different types of cities, the three parameters of extreme minimum temperature distribution of the urban station group are, without exception, higher than those of the reference station group. The urbanization of different types of cities all show a warming effect, with small-size cities have the most evident effects on extreme minimum temperatures.
    Liu W.,Coauthors, 2015: Extended reconstructed Sea surface temperature Version 4 (ERSST.v4): Part II. Parametric and structural uncertainty estimations. J.Climate, 28, 931-951.13805dbc660c09f00623555c09245545http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F272371313_Extended_Reconstructed_Sea_Surface_Temperature_Version_4_%28ERSST.v4%29_Part_II._Parametric_and_Structural_Uncertainty_Estimationshttp://www.researchgate.net/publication/272371313_Extended_Reconstructed_Sea_Surface_Temperature_Version_4_(ERSST.v4)_Part_II._Parametric_and_Structural_Uncertainty_Estimations
    Lugina K. M.,P. Y. Groisman,K. Y. Vinnikov,V. V. Koknaeva, and N. A. Speranskaya, 2006: Monthly surface air temperature time series area-averaged over the 30-degree latitudinal belts of the globe, 1881-2005. Trends: A Compendium of Data on Global Change. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Dept. Energy, Oak Ridge, Tenn., U.S.A.874b9bbaeaa910a4b7fe74a35bd58cbfhttp%3A%2F%2Fcdiac.esd.ornl.gov%2Ftrends%2Ftemp%2Flugina%2Fluginahttp://cdiac.esd.ornl.gov/trends/temp/lugina/luginaThe mean monthly and annual values of surface air temperature compiled by Lugina have been taken mainly from the , , and . These published records were supplemented with information from different national publications. In the original archive, after removal of station records believed to be nonhomogeneous or biased, 301 and 265 stations were used to determine the mean temperature for the Northern and Southern hemispheres, respectively. The new version of the station temperature archive (used for evaluation of the zonally-averaged temperatures) was created in 1995. The change to the archive was required because data from some stations became unavailable for analyses in the 1990s. During this process, special care was taken to secure homogeneity of zonally averaged time series. When a station (or a group of stations) stopped reporting, a "new" station (or group of stations) was selected in the same region, and its data for the past 50 years were collected and added to the archive. The processing (area-averaging) was organized in such a way that each time series from a new station spans the reference period (1951-1975) and the years thereafter. It was determined that the addition of the new stations had essentially no effect on the zonally-averaged values for the pre-1990 period.
    Masson-Delmotte V. M., and Coauthors, 2013: Information from paleoclimate archives. Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, T. F. Stocker et al., Eds. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.da927e38df4dfaf1ef4475a4dd16b14ehttp%3A%2F%2Fueaeprints.uea.ac.uk%2F37843%2Fhttp://ueaeprints.uea.ac.uk/37843/Cover photo: The Blue Marble western and eastern hemispheres. These images integrate land, ocean, sea ice and clouds into a visual representation of the earth's climate system. They are based on space-borne earth observation data from NASA's MODIS (MODerate
    Maury M. F.,1855: Wind and Current Charts. 7th ed., US Navy, Philadelphia.
    Menne M. J.,C. N. Williams Jr., and R. S. Vose, 2009: The U.S historical climatology network monthly temperature data, Version 2. Bull. Amer. Meteor. Soc., 90, 993-1007.10.1175/2008BAMS2613.153746d71-be32-471d-a57d-4b711a7293f37a80cede463313e4eb50f49b8beeeccchttp%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20093241932.htmlrefpaperuri:(cc325e9980d96706eec51033ba2211de)http://www.cabdirect.org/abstracts/20093241932.htmlAbstract In support of climate monitoring and assessments, the National Oceanic and Atmospheric Administration's (NOAA's) National Climatic Data Center has developed an improved version of the U.S. Historical Climatology Network temperature dataset (HCN version 2). In this paper, the HCN version 2 temperature data are described in detail, with a focus on the quality-assured data sources and the systematic bias adjustments. The bias adjustments are discussed in the context of their effect on U.S. temperature trends from the period 1895-2007 and in terms of the differences between version 2 and its widely used predecessor (now referred to as HCN version 1). Evidence suggests that the collective effect of changes in observation practice at U.S. HCN stations is systematic and of the same order of magnitude as the background climate signal. For this reason, bias adjustments are essential to reducing the uncertainty in U.S. climate trends. The largest biases in the HCN are shown to be associated with changes to the time of observation and with the widespread changeover from liquid-in-glass thermometers to the maximum-搈inimum temperature system (MMTS). With respect to HCN version 1, HCN version 2 trends in maximum temperatures are similar, while minimum temperature trends are somewhat smaller because of 1) an apparent overcorrection in HCN version 1 for the MMTS instrument change and 2) the systematic effect of undocumented station changes, which were not addressed in HCN version 1.
    Moberg A.,H. Alexandersson,H. Bergstr枚m, and P. D. Jones, 2003: Were southern Swedish summer temperatures before 1860 as warm as measured? Inter. J. Climatol., 23, 1495-1521.10.1002/joc.945b11121dd5f3dd0651ac3ea4c9d5bf68ahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.945%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1002/joc.945/citedbyAbstract Temperature series from Stockholm and Uppsala in southern Sweden indicate that summers from the mid-18th century until around 1860 were, on average, warmer than the 1961–90 mean. The station histories suggest that the early observations could have been positively biased, for example because of insufficient radiation protection. We investigate if independent support for warm summers in the early period can be obtained from other climate variables. Using stepwise multiple regression analysis we investigate nine potential predictor variables: six air circulation indices, precipitation, air pressure and cloud amount. Three of these variables—cloud amount (the most important one), meridional geostrophic wind, and air pressure—together explain 65% of the June–August temperature variance in the calibration period 1873–2000. Application of the regression relationship back to 1780 shows that the model is equally successful in predicting year-to-year temperature variability before 1873 as it is in the calibration period, whereas the low-frequency component is poorly reconstructed in the early period. This reduced skill is primarily due to poorer data quality of the predictor variables in the early period, in particular the cloud amount series. The observed decadal mean temperatures during 1780–1860 are found to be above the upper limit of a 95% confidence interval that accounts for uncertainties both in the regression relationship and in the cloud amount series. We conclude that the observed temperatures before around 1860 are, therefore, most likely positively biased. The size of this bias cannot be accurately determined from the evidence used here, but seems to be about 0.7–0.8°C for both stations. A comparison with long instrumental temperature series from central Europe suggests a slightly smaller bias (0.5–0.6°C). For more accurate assessment of the Stockholm and Uppsala temperatures, we recommend that extensive homogeneity testing of other long northern European temperature series are undertaken. Copyright 08 2003 Royal Meteorological Society
    Morice C. P.,J. J. Kennedy,N. A. Rayner, and P. D. Jones, 2012: Quantifying uncertainties in global and regional temperature change using an ensemble of observational estimates: the HadCRUT4 data set. J. Geophys. Res., 117, D08101, doi: 10.1029/2011JD017187.10.1029/2011JD0171878b1cc10538405cb9260ddc3ff5fdae8bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2011JD017187%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2011JD017187/abstractABSTRACT Recent developments in observational near-surface air temperature and sea-surface temperature analyses are combined to produce HadCRUT4, a new data set of global and regional temperature evolution from 1850 to the present. This includes the addition of newly digitized measurement data, both over land and sea, new sea-surface temperature bias adjustments and a more comprehensive error model for describing uncertainties in sea-surface temperature measurements. An ensemble approach has been adopted to better describe complex temporal and spatial interdependencies of measurement and bias uncertainties and to allow these correlated uncertainties to be taken into account in studies that are based upon HadCRUT4. Climate diagnostics computed from the gridded data set broadly agree with those of other global near-surface temperature analyses. Fitted linear trends in temperature anomalies are approximately 0.07ºC/decade from 1901 to 2010 and 0.17ºC/decade from 1979 to 2010 globally. Northern/southern hemispheric trends are 0.08/0.07ºC/decade over 1901 to 2010 and 0.24/0.10ºC/decade over 1979 to 2010. Linear trends in other prominent near-surface temperature analyses agree well with the range of trends computed from the HadCRUT4 ensemble members.
    Nicholls N.,R. Tapp,K. Burrows, and D. Richards, 1996: Historical thermometer exposures in Australia. Inter. J. Climatol., 16, 705-710.10.1002/(SICI)1097-0088(199606)16:63.0.CO;2-Sd675d74b1f3e105745820ae77ecdfa95http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F%28SICI%291097-0088%28199606%2916%3A6%3C705%3A%3AAID-JOC30%3E3.0.CO%3B2-S%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/(SICI)1097-0088(199606)16:6<705::AID-JOC30>3.0.CO;2-S/fullABSTRACT
    Parker D. E.,1994: Effects of changing exposure of thermometers at land stations. Inter. J. Climatol., 14, 1-31.10.1002/joc.33701401024f479578dd64bca5fc28e4e8e3fa0ad6http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.3370140102%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/joc.3370140102/abstractABSTRACT In view of the implications for the assessment of climatic changes since the mid-nineteenth century, systematic changes of exposure of thermometers at land stations are reviewed. Particular emphasis is laid on changes of exposure during the late nineteenth and early twentieth century when shelters often differed considerably from the Stevenson screens, and variants thereof, which have been prevalent during the past few decades. It is concluded that little overall bias in land surface air temperature has accumulated since the late nineteenth century: however, the earliest extratropical data may have been biased typically 0.2ºC warm in summer and by day, and similarly cold in winter and by night, relative to modern observations. Furthermore, there is likely to have been a warm bias in the tropics in the early twentieth century: this bias, implied by comparisons between Stevenson screens and the tropical sheds then in use, is confirmed by comparisons between coastal land surface air temperatures and nearby marine surface temperatures, and was probably of the order of 0.2ºC.
    Parker D. E.,2004: Climate: large-scale warming is not urban. Nature, 432, 290 pp.10.1038/432290a155490879b6c4e5ed55e5b5d5204df5b5ab141ebhttp%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM15549087http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM15549087Controversy has persisted over the influence of urban warming on reported large-scale surface-air temperature trends. Urban heat islands occur mainly at night and are reduced in windy conditions. Here we show that, globally, temperatures over land have risen as much on windy nights as on calm nights, indicating that the observed overall warming is not a consequence of urban development.
    Parker D. E.,2006: A demonstration that large-scale warming is not urban. J.Climate, 19, 2882-2895.10.1175/JCLI3730.158abc94d1013358056e53574fde1cab1http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F260301822_A_Demonstration_That_Large-Scale_Warming_Is_Not_Urbanhttp://www.researchgate.net/publication/260301822_A_Demonstration_That_Large-Scale_Warming_Is_Not_UrbanAbstract On the premise that urban heat islands are strongest in calm conditions but are largely absent in windy weather, daily minimum and maximum air temperatures for the period 1950-2000 at a worldwide selection of land stations are analyzed separately for windy and calm conditions, and the global and regional trends are compared. The trends in temperature are almost unaffected by this subsampling, indicating that urban development and other local or instrumental influences have contributed little overall to the observed warming trends. The trends of temperature averaged over the selected land stations worldwide are in close agreement with published trends based on much more complete networks, indicating that the smaller selection used here is sufficient for reliable sampling of global trends as well as interannual variations. A small tendency for windy days to have warmed more than other days in winter over Eurasia is the opposite of that expected from urbanization and is likely to be a consequence of atmospheric circulation changes.
    Parker D. E.,2010: Urban heat island effects on estimates of observed climate change. Wiley Interdisciplinary Reviews: Climate Change, 1(1), 123-133, doi: 10.1002/wcc.21.10.1002/wcc.210939c0b0d73e810cc8f3932823dff59bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fwcc.21%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/wcc.21/abstractAbstract Urban heat islands are a result of the physical properties of buildings and other structures, and the emission of heat by human activities. They are most pronounced on clear, calm nights; their strength depends also on the background geography and climate, and there are often cool islands in parks and less-developed areas. Some old city centers no longer show warming trends relative to rural neighbourhoods, because urban development has stabilised. This article reviews the effects that urban heat islands may have on estimates of global near-surface temperature trends. These effects have been reduced by avoiding or adjusting urban temperature measurements. Comparisons of windy weather with calm-weather air temperature trends for a worldwide set of observing sites suggest that global near-surface temperature trends have not been greatly affected by urban warming trends; this is supported by comparisons with marine surface temperatures. The use of dynamical-model-based reanalyses to estimate urban influences has been hindered by the heterogeneity of the data input to the reanalyses and by biases in the models. However, improvements in reanalyses are increasing their utility for assessing the surface air temperature record. High-resolution climate models and data on changing land use offer potential for future assessment of worldwide urban warming influences. The latest assessments of the likely magnitude of the residual urban trend in available global near-surface temperature records are summarized, along with the uncertainties of these residual trends. Copyright 2010 John Wiley & Sons, Ltd. For further resources related to this article, please visit the TODO: clickthrough URL WIREs website .
    Parker D. E.,2011: Recent land surface air temperature trends assessed using the 20th century reanalysis. J. Geophys. Res., 116, D20125, doi: 10.1029/2011JD016438.10.1029/2011JD016438eeb48399-b02f-43dc-b390-965da213ee70b3264c7f93639efa3fd8742327a146dchttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2011JD016438%2Fabstractrefpaperuri:(5177724de0ff8c88ef6d80b17abf40c3)http://onlinelibrary.wiley.com/doi/10.1029/2011JD016438/abstract[1] Land surface air temperature trends observed during 1979&ndash;2008 are compared with those simulated by the 20th Century Reanalysis that is driven only by observed sea surface temperatures and sea ice, atmospheric CO 2 concentrations, solar and volcanic forcings, and surface pressure data. On a global annual average, the 20th Century Reanalysis simulates a little more than 80% of the observed trend, but with substantial regional and seasonal variations. The remainder of the trend may be ascribed tentatively to land use changes, aerosol increases and decreases, and changes in minor greenhouse gases not accounted for in the 20th Century Reanalysis.
    Parker D. E.,P. Jones,T. C. Peterson, and J. Kennedy, 2009: Comment on -淯nresolved issues with the assessment of multidecadal global land surface temperature trends- by Roger A. Pielke Sr. et al. J. Geophys. Res., 114, D05104, doi: 10.1029/ 2008JD010450.10.1029/2008JD0104502f27fbc8a7abe569e76eec79d4770ef4http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2008JD010450%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2008JD010450/fullFirst page of article
    Peterson T. C.,T. W. Owen, 2005: Urban heat island assessment: metadata are important. J.Climate, 18, 2637-2646.10.1175/JCLI3431.1ab700801bed3eb8b355d3d8b750df6e9http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F253664986_Urban_Heat_Island_Assessment_Metadata_Are_Importanthttp://www.researchgate.net/publication/253664986_Urban_Heat_Island_Assessment_Metadata_Are_ImportantAbstract Urban heat island (UHI) analyses for the conterminous United States were performed using three different forms of metadata: nightlights-derived metadata, map-based metadata, and gridded U.S. Census Bureau population metadata. The results indicated that metadata do matter. Whether a UHI signal was found depended on the metadata used. One of the reasons is that the UHI signal is very weak. For example, population was able to explain at most only a few percent of the variance in temperature between stations. The nightlights metadata tended to classify lower population stations as rural compared to map-based metadata while the map-based metadata urban stations had, on average, higher populations than urban nightlights. Analysis with gridded population metadata indicated that statistically significant urban heat islands could be found even when quite urban stations were classified as rural, indicating that the primary signal was coming from the relatively high population sites. If 鈭30% of the highest population stations were removed from the analysis, no statistically significant urban heat island was detected. The implications of this work on U.S. climate change analyses is that, if the highest population stations are avoided (populations above 30-000 within 6 km), the analysis should not be expected to be contaminated by UHIs. However, comparison between U.S. Historical Climatology Network (HCN) time series from the full dataset and a subset excluding the high population sites indicated that the UHI contamination from the high population stations accounted for very little of the recent warming.
    Poli, P.,Coauthors, 2013: The data assimilation system and initial performance evaluation of the ECMWF pilot reanalysis of the 20th-century assimilating surface observations only (ERA-20C). ERA Report Series, 14 pp.
    Quayle R. G.,D. R. Easterling,T. R. Karl, and P. Y. Hughes, 1991: Effects of recent thermometer changes in the cooperative station network. Bull. Amer. Meteor. Soc., 72, 1718-1723.10.1175/1520-0477(1991)072<1718:EORTCI>2.0.CO;2df99dee039b7cf9d8de18bd29f09be4bhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F234281066_Effects_of_Recent_Thermometer_Changes_in_the_Cooperative_Station_Networkhttp://www.researchgate.net/publication/234281066_Effects_of_Recent_Thermometer_Changes_in_the_Cooperative_Station_NetworkAbstract During the past five years, the National Weather Service (NWS) has replaced over half of its liquid-in-glass maximum and minimum thermometers in wooden Cotton Region Shelters (CRSS) with thermistor-based Maximum-Minimum Temperature Systems (MMTSS) housed in smaller plastic shelters. Analyses of data from 424 (of the 3300) MMTS stations and 675 CRS stations show that a mean daily minimum temperature change of roughly +0.3°C, a mean daily maximum temperature change of 610.4°C, and a change in average temperature of 610.1°C were introduced as a result of the new instrumentation. The change of 610.7°C in daily temperature range is particularly significant for climate change studies that use this element as an independent variable. Although troublesome for climatologists, there is reason to believe that this change (relative to older records) represents an improvement in absolute accuracy. The bias appears to be rather sharp and well defined. Since the National Climatic Data Center (NCDC) station history database contains records of instrumentation, adjustments for this bias can be readily applied, and we are reasonably confident that the corrections we have developed can be used to produce homogeneous time series of area-average temperature.
    Ren G. Y.,Y. Q. Zhou,Z. Y. Chu,J. X. Zhou,A. Y. Zhang,J. Guo, and X. F. Liu, 2008: Urbanization effects on observed surface air temperature trends in North China. J.Climate, 21, 1333-1348.10.1175/2007JCLI1348.1d9c3b51bd7ef6f4cac7e24d8011756fchttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F249611104_Urbanization_Effects_on_Observed_Surface_Air_Temperature_Trends_in_North_Chinahttp://www.researchgate.net/publication/249611104_Urbanization_Effects_on_Observed_Surface_Air_Temperature_Trends_in_North_ChinaABSTRACT A dataset of 282 meteorological stations including all of the ordinary and national basic/reference surface stations of north China is used to analyze the urbanization effect on surface air temperature trends. These stations are classified into rural, small city, medium city, large city, and metropolis based on the updated information of total population and specific station locations. The significance of urban warming effects on regional average temperature trends is estimated using monthly mean temperature series of the station group datasets, which undergo inhomogeneity adjustment. The authors found that the largest effect of urbanization on annual mean surface air temperature trends occurs for the large-city station group, with the urban warming being 0.16ºC/10yr, and the effect is the smallest for the small-city station group with urban warming being only 0.07ºC/10yr. A similar assessment is made for the dataset of national basic/reference stations, which has been widely used in regional climate change analyses in China. The results indicate that the regional average annual mean temperature series, as calculated using the data from the national basic/reference stations, is significantly impacted by urban warming, and the trend of urban warming is estimated to be 0.11ºC/10yr. The contribution of urban warming to total annual mean surface air temperature change as estimated with the national basic/reference station dataset reaches 37.9%. It is therefore obvious that, in the current regional average surface air temperature series in north China, or probably in the country as a whole, there still remain large effects from urban warming. The urban warming bias for the regional average temperature anomaly series is corrected. After that, the increasing rate of the regional annual mean temperature is brought down from 0.29ºC/10yr to 0.18ºC/10yr, and the total change in temperature approaches 0.72ºC for the period analyzed.
    Rennie, J.,Coauthors, 2014: The international surface temperature initiative global land surface databank: Monthly temperature data release description and methods. Geoscience Data Journal, 1, 75-102, doi: 10.1002/gdj3.8.10.1002/gdj3.8ff9b7fcecc8f3c56a8f60d35d16ad8aahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fgdj3.8%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/gdj3.8/fullAbstract Top of page Abstract Dataset Introduction and rationale 1Databank architecture 2Merging methodology 3Results of stage 3 dataset 4Data access and version control 5Concluding remarks and outlook Acknowledgements References AppendixA AppendixB Described herein is the first version release of monthly temperature holdings of a new Global Land Surface Meteorological Databank. Organized under the auspices of the International Surface Temperature Initiative (ISTI), an international group of scientists have spent threeyears collating and merging data from numerous sources to create a merged holding. This release in its recommended form consists of over 30000 individual station records, some of which extend over the past 300years. This article describes the sources, the chosen merge methodology, and the resulting databank characteristics. Several variants of the databank have also been released that reflect the structural uncertainty in merging datasets. Variants differ in, for example, the order in which sources are considered and the degree of congruence required in station geolocation for consideration as a merged or unique record. Also described is a version control protocol that will be applied in the event of updates. Future updates are envisaged with the addition of new data sources, and with changes in processing, where public feedback is always welcomed. Major updates, when necessary, will always be accompanied by a new journal paper. This databank release forms the foundation for the construction of new global land surface air temperature analyses by the global research community and their assessment by the ISTI's benchmarking and assessment working group.
    Rohde R.,Coauthors, 2013a: A new estimate of the average earth surface land temperature spanning 1753 to 2011. Geoinfor Geostat: An Overview, 1, doi: 10.4172/2327-4581. 1000101.10.4172/2327-4581.1000101002ca71ea56cdb6e3be2de2953212c96http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F269552143_A_New_Estimate_of_the_Average_Earth_Surface_Land_Temperature_Spanning_1753_to_2011http://www.researchgate.net/publication/269552143_A_New_Estimate_of_the_Average_Earth_Surface_Land_Temperature_Spanning_1753_to_2011Abstract We report an estimate of the Earth's average land surface temperature for the period 1753 to 2011. To address issues of potential station selection bias, we used a larger sampling of stations than had prior studies. For the period post 1880, our estimate is
    Rohde R.,Coauthors, 2013b: Berkeley earth temperature averaging process. Geoinfor Geostat: An Overview, 1, doi: 10.4172/ gigs.1000103.
    Simmons A. J.,K. M. Willett,P. D. Jones,P. W. Thorne, and D. P. Dee, 2010: Low-frequency variations in surface atmospheric humidity, temperature, and precipitation: inferences from reanalyses and monthly gridded observational data sets. J. Geophys. Res., 115, D01110, doi: 10.1029/2009JD012442.10.1029/2009JD012442a9f80d20-31cc-4cd5-975d-62a2657642a62196c3170ab15cad72d3995b1884c887http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2009JD012442%2Ffullrefpaperuri:(36a8c948bd48616f3204aefda29e0984)http://onlinelibrary.wiley.com/doi/10.1029/2009JD012442/fullABSTRACT Evidence is presented of a reduction in relative humidity over low-latitude and midlatitude land areas over a period of about 10 years leading up to 2008, based on monthly anomalies in surface air temperature and humidity from comprehensive European Centre for Medium-Range Weather Forecasts reanalyses (ERA-40 and ERA-Interim) and from Climatic Research Unit and Hadley Centre analyses of monthly station temperature data (CRUTEM3) and synoptic humidity observations (HadCRUH). The data sets agree well for both temperature and humidity variations for periods and places of overlap, although the average warming over land is larger for the fully sampled ERA data than for the spatially and temporally incomplete CRUTEM3 data. Near-surface specific humidity varies similarly over land and sea, suggesting that the recent reduction in relative humidity over land may be due to limited moisture supply from the oceans, where evaporation has been limited by sea surface temperatures that have not risen in concert with temperatures over land. Continental precipitation from the reanalyses is compared with a new gauge-based Global Precipitation Climatology Centre (GPCC) data set, with the combined gauge and satellite products of the Global Precipitation Climatology Project (GPCP) and the Climate Prediction Center (CPC), Merged Analysis of Precipitation (CMAP), and with CPC's independent gauge analysis of precipitation over land (PREC/L). The reanalyses agree best with the new GPCC and latest GPCP data sets, with ERA-Interim significantly better than ERA-40 at capturing monthly variability. Shifts over time in the differences among the precipitation data sets make it difficult to assess their longer-term variations and any link with longer-term variations in humidity.
    Smith T. M.,R. W., and Reynolds, 2005: A global merged land and sea surface temperature reconstruction based on historical observations (1880-1997). J Climate, 18, 2021-2036.71163d12-eb8b-4005-9721-e8ffba1333a3e662deaa50a49ec7153bc67a9cd75cd2http%3A%2F%2Fwww.readcube.com%2Farticles%2F10.1175%2FJCLI3362.1refpaperuri:(129707b446e6d030756a2a004543f764)http://www.readcube.com/articles/10.1175/JCLI3362.1
    Smith T. M.,R. W. Reynolds,T. C. Peterson, and J. Lawrimore, 2008: Improvements to NOAA-檚 historical merged Land-Ocean surface temperature analysis (1880-2006). J.Climate, 21, 2283-2293.
    Thompson D. W. J.,J. J. Kennedy,J. M. Wallace, and P. D. Jones, 2008: A large discontinuity in the mid-twentieth century in observed global-mean surface temperature. Nature, 453, 646-649.10.1038/nature0698218509442b3b319cfae56286172c9db921cee2f6dhttp%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM18509442http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM18509442ABSTRACT Data sets used to monitor the Earth's climate indicate that the surface of the Earth warmed from approximately 1910 to 1940, cooled slightly from approximately 1940 to 1970, and then warmed markedly from approximately 1970 onward. The weak cooling apparent in the middle part of the century has been interpreted in the context of a variety of physical factors, such as atmosphere-ocean interactions and anthropogenic emissions of sulphate aerosols. Here we call attention to a previously overlooked discontinuity in the record at 1945, which is a prominent feature of the cooling trend in the mid-twentieth century. The discontinuity is evident in published versions of the global-mean temperature time series, but stands out more clearly after the data are filtered for the effects of internal climate variability. We argue that the abrupt temperature drop of approximately 0.3 degrees C in 1945 is the apparent result of uncorrected instrumental biases in the sea surface temperature record. Corrections for the discontinuity are expected to alter the character of mid-twentieth century temperature variability but not estimates of the century-long trend in global-mean temperatures.
    Thompson D. W. J.,J. M. Wallace,P. D. Jones, and J. J. Kennedy, 2009: Identifying signatures of natural climate variability in time series of global-mean surface temperature: Methodology and insights. J.Climate, 22, 6120-6141.10.1175/2009JCLI3089.166cdaccecfa112733e2a699c42918016http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20103007997.htmlhttp://www.cabdirect.org/abstracts/20103007997.htmlAbstract Global-mean surface temperature is affected by both natural variability and anthropogenic forcing. This study is concerned with identifying and removing from global-mean temperatures the signatures of natural climate variability over the period January 1900–March 2009. A series of simple, physically based methodologies are developed and applied to isolate the climate impacts of three known sources of natural variability: the El Ni09o–Southern Oscillation (ENSO), variations in the advection of marine air masses over the high-latitude continents during winter, and aerosols injected into the stratosphere by explosive volcanic eruptions. After the effects of ENSO and high-latitude temperature advection are removed from the global-mean temperature record, the signatures of volcanic eruptions and changes in instrumentation become more clearly apparent. After the volcanic eruptions are subsequently filtered from the record, the residual time series reveals a nearly monotonic global warming pattern since 651950. The results also reveal coupling between the land and ocean areas on the interannual time scale that transcends the effects of ENSO and volcanic eruptions. Globally averaged land and ocean temperatures are most strongly correlated when ocean leads land by 652–3 months. These coupled fluctuations exhibit a complicated spatial signature with largest-amplitude sea surface temperature perturbations over the Atlantic Ocean.
    Trenberth K.E.,Coauthors, 2007: Observations: surface and atmospheric climate change. Climate Change 2007: The Physical Science Basis. Contribution of Working Group 1 to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, S. D. Solomon et al., Eds. Cambridge University Press, 235-336.187d9a7d7fd5d58cd42bcdc3baaeee7chttp%3A%2F%2Fwww.scientists.org.nz%2Fpublications%2Fview%2F175http://www.scientists.org.nz/publications/view/175
    Trewin B.,2010: Exposure, instrumentation, and observing practice effects on land temperature measurements. Wiley Interdisciplinary Reviews: Climate Change, 1, 490-506, doi: 10.1002/wcc.46, 2010.10.1002/wcc.46eb3420099cad2e96e6b77242879a18f8http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fwcc.46%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/wcc.46/abstractAbstract To monitor climate change adequately and determine the extent to which anthropogenic influences are contributing to observed climate change, it is critical to have land temperature data of a high standard. In particular, it is important to have temperature data whose changes reflect changes in the climate and not changes in other circumstances under which the temperatures were taken. There are numerous factors that can affect land temperature records. Among the most common are changes in instrumentation, changes in local site condition in situ (through urbanization or for other reasons), site relocations, and changes in observing practices. All have the potential, if uncorrected, to have impacts on temperature records at individual locations similar to or greater than the observed century-scale global warming trend. A number of techniques exist to identify these influences and correct data to take them into account. These have been applied in various ways in climate change analyses and in major data sets used for the assessment of long-term climate change. These techniques are not perfect and numerous uncertainties remain, especially with respect to daily and sub-daily temperature data. Copyright 2010 John Wiley & Sons, Ltd. For further resources related to this article, please visit the TODO: clickthrough URL WIREs website
    Venema, V. K. C.,Coauthors, 2012: Benchmarking homogenization algorithms for monthly data. Climates of the Past, 8, 89-115.10.1063/1.4819690d266a149-91a3-4cf5-80ba-14b064cd19d65ccd7e37c7a115a7c1565ddbb19733b6http%3A%2F%2Fscitation.aip.org%2Fcontent%2Faip%2Fproceeding%2Faipcp%2F10.1063%2F1.4819690refpaperuri:(7477aa6100c003faa6daa62dc602f7fa)http://scitation.aip.org/content/aip/proceeding/aipcp/10.1063/1.4819690The COST (European Cooperation in Science and Technology) Action ES0601: advances in homogenization methods of climate series: an integrated approach (HOME) has executed a blind intercomparison and validation study for monthly homogenization algorithms. Time series of monthly temperature and precipitation were evaluated because of their importance for climate studies and because they represent two important types of statistics (additive and multiplicative). The algorithms were validated against a realistic benchmark dataset. The benchmark contains real inhomogeneous data as well as simulated data with inserted inhomogeneities. Random independent break-type inhomogeneities with normally distributed breakpoint sizes were added to the simulated datasets. To approximate real world conditions, breaks were introduced that occur simultaneously in multiple station series within a simulated network of station data. The simulated time series also contained outliers, missing data periods and local station trends. Further, a stochastic nonlinear global (network-wide) trend was added. Participants provided 25 separate homogenized contributions as part of the blind study. After the deadline at which details of the imposed inhomogeneities were revealed, 22 additional solutions were submitted. These homogenized datasets were assessed by a number of performance metrics including (i) the centered root mean square error relative to the true homogeneous value at various averaging scales, (ii) the error in linear trend estimates and (iii) traditional contingency skill scores. The metrics were computed both using the individual station series as well as the network average regional series. The performance of the contributions depends significantly on the error metric considered. Contingency scores by themselves are not very informative. Although relative homogenization algorithms typically improve the homogeneity of temperature data, only the best ones improve precipitation data. Training the users on homogenization software was found to be very important. Moreover, stateof- the-art relative homogenization algorithms developed to work with an inhomogeneous reference are shown to perform best. The study showed that automatic algorithms can perform as well as manual ones.
    Vose, R. S.,Coauthors, 2012: NOAA-檚 merged Land-Ocean surface temperature analysis. Bull. Amer. Meteor. Soc., 93, 1677-1685, doi: 10.1175/BAMS-D-11-00241.1.10.1175/BAMS-D-11-00241.148250f9e-8fd0-4240-95f9-80b18bb23d11a871f494927b97ada299e482d296dab1http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F273920081_NOAA%27s_Merged_LandOcean_Surface_Temperature_Analysisrefpaperuri:(dbc44c65ee1c9ace06b0727517ae2bee)http://www.researchgate.net/publication/273920081_NOAA's_Merged_LandOcean_Surface_Temperature_AnalysisThis paper describes the new release of the Merged Land–Ocean Surface Temperature analysis (MLOST version 3.5), which is used in operational monitoring and climate assessment activities by the NOAA National Climatic Data Center. The primary motivation for the latest version is the inclusion of a new land dataset that has several major improvements, including a more elaborate approach for addressing changes in station location, instrumentation, and siting conditions. The new version is broadly consistent with previous global analyses, exhibiting a trend of 0.076°C decade 611 since 1901, 0.162°C decade 611 since 1979, and widespread warming in both time periods. In general, the new release exhibits only modest differences with its predecessor, the most obvious being very slightly more warming at the global scale (0.004°C decade 611 since 1901) and slightly different trend patterns over the terrestrial surface.
    Wang F.,Q. S. Ge,S. W. Wang,Q. X. Li, and P. D. Jones, 2015: A new estimation of urbanization-檚 contribution to the warming trend in China. J.Climate, 28, 8923-8938, doi: 10.1175/JCLI-D-14-00427.1.10.1175/JCLI-D-14-00427.1abbdbcab-e3f4-4fbc-aa52-064c5130ff9ad5ba8b04955e409207c2cf9b9504c1d8http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F282471171_A_New_Estimation_of_Urbanizations_Contribution_to_the_Warming_Trend_in_Chinahttp://www.researchgate.net/publication/282471171_A_New_Estimation_of_Urbanizations_Contribution_to_the_Warming_Trend_in_ChinaAbstract The extent to which an urbanization effect has contributed to climate warming is under debate in China. Some previous studies have shown that the urban heat island (UHI) contribution to national warming was substantial (10%-40%). However, by considering the spatial scale of urbanization effects, this study indicates that the UHI contribution is negligible (less than 1%). Urban areas constitute only 0.7% of the whole of China. According to the proportions of urban and rural areas used in this study, the weighted urban and rural temperature averages reduced the estimated total warming trend and also reduced the estimated urban effects. Conversely, if all stations were arithmetically averaged, that is, without weighting, the total warming trend and urban effects will be overestimated as in previous studies because there are more urban stations than rural stations in China. Moreover, the urban station proportion (68%) is much higher than the urban area proportion (0.7%).
    Wang J.,Z. W. Yan,P. D. Jones, and J. J. Xia, 2013: On -渙bservation minus reanalysis-� method: A view from multidecadal variability. J. Geophys. Res., 118, 7450-7458, doi: 10.1002/ jgrd.50574.10.1002/jgrd.5057487060ff2-9383-49c4-bd98-eb66aea474175196c519d7bc6792a0945febec8ae144http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjgrd.50574%2Ffullrefpaperuri:(ce6689851f53e525ef15d808a08a1045)http://onlinelibrary.wiley.com/doi/10.1002/jgrd.50574/fullAbstract Top of page Abstract 1Introduction 2Data and Methods 3Results 4Discussion and Summary Acknowledgments References [1] The observation minus reanalysis (OMR) method is widely used to investigate the impact of urbanization and land use change on climate. Here we present the OMR trends for the periods of 1979–1998 and 1989–2008 in eastern China, which appear inconsistent for the regions experiencing rapid urbanization during recent decades. Using Ensemble Empirical Mode Decomposition, we extract the secular trend and multidecadal variability (MDV) from the temperature observations at stations and the corresponding reanalysis data for the last century and find that, in general, MDV in the reanalysis data is weaker than that in the station observations. This systematic difference considerably modulates the magnitude of the OMR trends during different periods, leading to inconsistent estimates of the impact of urbanization. After MDV adjustment, the OMR trends for Beijing and Shanghai are consistent for the different periods, about 0.04°C–0.1°C/decade, much smaller than some previous estimates. We caution those using OMR methods to estimate the effect of urbanization and also for those using reanalysis data for a limited period in studies of this kind.
    Wickham, C.,Coauthors, 2013: Influence of urban heating on the global temperature land average using rural sites identified from MODIS classifications. Geoinfor Geostat: An Overview, 1, doi: 10.4172/2327-4581.1000104.10.1029/2011JD01682438a72daa5d526ed358f68be08310dbffhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F255564622_Influence_of_urban_heating_on_the_global_temperature_land_average_using_rural_sites_identified_from_MODIS_classificationshttp://www.researchgate.net/publication/255564622_Influence_of_urban_heating_on_the_global_temperature_land_average_using_rural_sites_identified_from_MODIS_classificationsThe effect of urban heating on estimates of global average land surface temperature is studied by applying an urban-rural classification based on MODIS satellite data to the Berkeley Earth temperature dataset compilation of 39,028 sites from 10 different publicly available sources. We compare the distribution of linear temperature trends for these sites to the distribution for a rural subset of 16,132 sites chosen to be distant from all MODISidentified urban areas. While the trend distri
    Wilby R. L.,P. D. Jones, and D. H. Lister, 2011: Decadal variations in the nocturnal heat island of London. Weather, 66, 59-64.10.1002/wea.679b82d7bb4d6698736943e0e1f92ba7352http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fwea.679%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1002/wea.679/pdfFirst page of article
    Woodruff, S. D.,Coauthors, 2011: ICOADS release 2.5: extensions and enhancements to the surface marine meteorological archive . Inter. J.Climatol., 31, 951-967, doi: 10.1002/joc.2103.10.1002/joc.2103ecac71db0999fcac5002bf67a0256968http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.2103%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/joc.2103/fullAbstract Top of page Abstract 1.Introduction 2.The national and international contributing community 3.Release 2.5 4.Data characteristics and unresolved issues 5.Data and product access 6.Community data products derived from ICOADS observations 7.Future plans 8.Conclusions Acknowledgements References Release 2.5 of the International Comprehensive Ocean-Atmosphere Data Set (ICOADS) is a major update (covering 1662–2007) of the world's most extensive surface marine meteorological data collection. Building on extensive national and international partnerships, many new and improved contributing datasets have been processed into a uniform format and combined with the previous Release 2.4. The new data range from early non-instrumental ship observations to measurements initiated in the twentieth century from buoys and other automated platform types. Improvements to existing data include replacing preliminary Global Telecommunication System (GTS) receipts with more reliable, delayed mode reports for post-1997 data, and in the processing and quality control (QC) of humidity observations. Over the entire period of record, spatial and temporal coverage has been enriched and data and metadata quality has been improved. Along with the observations, now updated monthly in near real time, Release 2.5 includes quality-controlled monthly summary products for 2° latitude × 2° longitude (since 1800) and 1° × 1° boxes (since 1960), together with multiple options for access to the data and products. The measured and estimated data in Release 2.5 are subject to many technical changes, multiple archive sources, and historical events throughout the more than three-century record. Some of these data characteristics are highlighted, including known unresolved errors and inhomogeneities, which may impact climate and other research applications. Anticipated future directions for ICOADS aim to continue adding scientific value to the observations, products, and metadata, as well as strengthen the cooperative enterprise through expanded linkages to international initiatives and organisations. Copyright 08 2010 Royal Meteorological Society
    Xu W. H.,Q. X. Li,X. L. Wang,S. Yang,L. J. Cao, and Y. Feng, 2013: Homogenization of Chinese daily surface air temperatures and analysis of trends in the extreme temperature indices. J. Geophys. Res., 118, 9708-9720, doi: 10.1002/jgrd. 50791.10.1002/jgrd.50791e62008e7bfb1011d03e0987e965d5e54http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjgrd.50791%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/jgrd.50791/fullAbstract Top of page Abstract Introduction Data and Methods Statistics of Detected Change Points and the Related Causes Inhomogeneity Impacts on Estimation of Long-Term Trends Trends in Extreme Temperature Indices Derived From the Homogenized Data Summary and Discussion Acknowledgments References [1] This study first homogenizes time series of daily maximum and minimum temperatures recorded at 825 stations in China over the period from 1951 to 2010, using both metadata and the penalized maximum t test with the first-order autocorrelation being accounted for to detect change points and using the quantile-matching algorithm to adjust the data time series to diminish discontinuities. Station relocation was found to be the main cause for discontinuities, followed by station automation. The effects of discontinuities on estimation of long-term trends in the annual mean and extreme indices of temperature are illustrated. The data homogenization is shown to have improved the spatial consistency of estimated trends. Using the homogenized daily minimum and daily maximum temperature data, this study also analyzes trends in extreme temperature indices. The results show that the vast majority (85%-90%) of the 825 sites have experienced significantly more warm nights and less cold nights since 1951. There have also been more warm days and less cold days since 1951, although these trends are less extensive. About 62% of the 825 sites were found to have experienced significantly more warm days and about 50% significantly less cold days. None of the 825 sites were found to have significantly more cold nights/days or less warm nights/days. These indicate that the warming is stronger in nighttime than in daytime and stronger in winter than in summer. Thus, the diurnal temperature range was found to have significantly decreased at 49% of the 825 sites, with significant increases being identified only at 3% of these sites.
    Zhao P.,P. D. Jones,L. J. Cao,Z. W. Yan,S. Y. Zha,Y. N. Zhu,Y. Yu, and G. L. Tang, 2014: Trend of surface air temperature in eastern china and associated large-scale climate variability over the last 100 years. J. Climate. 27, 4693-4703, doi: 10.1175/JCLI-D-13-00397.1.10.1175/JCLI-D-13-00397.1b5e6b4c65e2fdc58de2f866511ea0866http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F284437417_Trend_of_Surface_Air_Temperature_in_Eastern_China_and_Associated_Large-Scale_Climate_Variability_over_the_Last_100_Yearshttp://www.researchgate.net/publication/284437417_Trend_of_Surface_Air_Temperature_in_Eastern_China_and_Associated_Large-Scale_Climate_Variability_over_the_Last_100_YearsAbstract Using the reconstructed continuous and homogenized surface air temperature (SAT) series for 16 cities across eastern China (where the greatest industrial developments in China have taken place) back to the nineteenth century, the authors examine linear trends of SAT. The regional-mean SAT over eastern China shows a warming trend of 1.52°C (100 yr) 611 during 1909–2010. It mainly occurred in the past 4 decades and this agrees well with the variability in another SAT series developed from a much denser station network (over 400 sites) across this part of China since 1951. This study collects population data for 245 sites (from these 400+ locations) and split these into five equally sized groups based on population size. Comparison of these five groups across different durations from 30 to 60 yr in length indicates that differences in population only account for between 9% and 24% of the warming since 1951. To show that a larger urbanization impact is very unlikely, the study additionally determines how much can be explained by some large-scale climate indices. Anomalies of large-scale climate indices such as the tropical Indian Ocean SST and the Siberian atmospheric circulation systems account for at least 80% of the total warming trends.
    Zhou L. M.,R. E. Dickinson,Y. H. Tian,J. Y. Fang,Q. X. Li,R. K. Kaufmann,C. J. Tucker, and R. B. Myneni, 2004: Evidence for a significant urbanization effect on climate in China. Proceedings of the National Academy of Sciences of the United States of America, 101, 9540-9544.10.1073/pnas.0400357101152054808321ce035d93509b2902de2311b05540http%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM15205480http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM15205480China has experienced rapid urbanization and dramatic economic growth since its reform process started in late 1978. In this article, we present evidence for a significant urbanization effect on climate based on analysis of impacts of land-use changes on surface temperature in southeast China, where rapid urbanization has occurred. Our estimated warming of mean surface temperature of 0.05 degrees C per decade attributable to urbanization is much larger than previous estimates for other periods and locations. The spatial pattern and magnitude of our estimate are consistent with those of urbanization characterized by changes in the percentage of urban population and in satellite-measured greenness.
  • [1] HAN Zuoqiang, YAN Zhongwei*, LI Zhen, LIU Weidong, and WANG Yingchun, 2014: Impact of Urbanization on Low-Temperature Precipitation in Beijing during 19602008, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 48-56.  doi: 10.1007/s00376-013-2211-3
    [2] YAN Li, WANG Panxing, YU Yongqiang, LI Lijuan, WANG Bin, 2010: Potential Predictability of Sea Surface Temperature in a Coupled Ocean--Atmosphere GCM, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 921-936.  doi: 10.1007/s00376-009-9062-y
    [3] Xiang LI, Tiejun LING, Yunfei ZHANG, Qian ZHOU, 2018: A 31-year Global Diurnal Sea Surface Temperature Dataset Created by an Ocean Mixed-Layer Model, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1443-1454.  doi: 10.1007/s00376-018-8016-7
    [4] LIN Pengfei, LIU Hailong, YU Yongqiang, ZHANG Xuehong, 2011: Response of Sea Surface Temperature to Chlorophyll-a Concentration in the Tropical Pacific: Annual Mean, Seasonal Cycle, and Interannual Variability, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 492-510.  doi: 10.1007/s00376-010-0015-2
    [5] Kui LIU, Lian-Tong ZHOU, Zhibiao WANG, Yong LIU, 2023: Interdecadal Enhancement in the Relationship between the Western North Pacific Summer Monsoon and Sea Surface Temperature in the Tropical Central-Western Pacific after the Early 1990s, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1766-1782.  doi: 10.1007/s00376-023-2200-0
    [6] Tianbao XU, Zhicong YIN, Xiaoqing MA, Yanyan HUANG, Huijun WANG, 2023: Hybrid Seasonal Prediction of Meridional Temperature Gradient Associated with “Warm Arctic–Cold Eurasia”, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1649-1661.  doi: 10.1007/s00376-023-2226-3
    [7] LIN Pengfei, LIU Hailong, ZHANG Xuehong, 2007: Sensitivity of the Upper Ocean Temperature and Circulation in the Equatorial Pacific to Solar Radiation Penetration Due to Phytoplankton, ADVANCES IN ATMOSPHERIC SCIENCES, 24, 765-780.  doi: 10.1007/s00376-007-0765-7
    [8] Yujie WU, Wansuo DUAN, 2018: Impact of SST Anomaly Events over the Kuroshio-Oyashio Extension on the "Summer Prediction Barrier", ADVANCES IN ATMOSPHERIC SCIENCES, 35, 397-409.  doi: 10.1007/s00376-017-6322-0
    [9] Ben TIAN, Hong-Li REN, 2022: Diagnosing SST Error Growth during ENSO Developing Phase in the BCC_CSM1.1(m) Prediction System, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 427-442.  doi: 10.1007/s00376-021-1189-5
    [10] LI Shuanglin, CHEN Xiaoting, 2014: Quantifying the Response Strength of the Southern Stratospheric Polar Vortex to Indian Ocean Warming in Austral Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 492-503.  doi: 10.1007/s00376-013-2322-x
    [11] Meng YAN, Johnny C. L. CHAN, Kun ZHAO, 2020: Impacts of Urbanization on the Precipitation Characteristics in Guangdong Province, China, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 696-706.  doi: 10.1007/s00376-020-9218-3
    [12] WANG Linlin, GAO Zhiqiu, MIAO Shiguang, GUO Xiaofeng, SUN Ting, Maofeng LIU, Dan LI, 2015: Contrasting Characteristics of the Surface Energy Balance between the Urban and Rural Areas of Beijing, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 505-514.  doi: 10.1007/s00376-014-3222-4
    [13] WANG Xuemei, CHEN Fei, WU Zhiyong, ZHANG Meigen, Mukul TEWARI, Alex GUENTHER, Christine WIEDINMYER, 2009: Impacts of Weather Conditions Modified by Urban Expansion on Surface Ozone: Comparison between the Pearl River Delta and Yangtze River Delta Regions, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 962-972.  doi: 10.1007/s00376-009-8001-2
    [14] Ping LIANG, Yihui DING, 2017: The Long-term Variation of Extreme Heavy Precipitation and Its Link to Urbanization Effects in Shanghai during 1916-2014, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 321-334.  doi: 10.1007/s00376-016-6120-0
    [15] LI Zhen, YAN Zhongwei, TU Kai, WU Hongyi, 2015: Changes of Precipitation and Extremes and the Possible Effect of Urbanization in the Beijing Metropolitan Region during 1960-2012 Based on Homogenized Observations, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1173-1185.  doi: 10.1007/s00376-015-4257-x
    [16] Yun QIAN, TC CHAKRABORTY, Jianfeng LI, Dan LI, Cenlin HE, Chandan SARANGI, Fei CHEN, Xuchao YANG, L. Ruby LEUNG, 2022: Urbanization Impact on Regional Climate and Extreme Weather: Current Understanding, Uncertainties, and Future Research Directions, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 819-860.  doi: 10.1007/s00376-021-1371-9
    [17] Yali LUO, Jiahua ZHANG, Miao YU, Xudong LIANG, Rudi XIA, Yanyu GAO, Xiaoyu GAO, Jinfang YIN, 2023: On the Influences of Urbanization on the Extreme Rainfall over Zhengzhou on 20 July 2021: A Convection-Permitting Ensemble Modeling Study, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 393-409.  doi: 10.1007/s00376-022-2048-8
    [18] Tang Youhua, Miao Manqian, 1998: Numerical Studies on Urban Heat Island Associated with Urbanization in Yangtze Delta Region, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 393-403.  doi: 10.1007/s00376-998-0009-5
    [19] LIU Juan, WANG Bin, LIU Hailong, YU Yongqiang, 2008: A New Global Four-Dimensional Variational Ocean Data Assimilation System and Its Application, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 680-691.  doi: 10.1007/s00376-008-0680-6
    [20] Shuangmei MA, Congwen ZHU, Juan LIU, 2020: Combined Impacts of Warm Central Equatorial Pacific Sea Surface Temperatures and Anthropogenic Warming on the 2019 Severe Drought in East China, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1149-1163.  doi: 10.1007/s00376-020-0077-8

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 31 August 2015
Manuscript accepted: 09 November 2015
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

The Reliability of Global and Hemispheric Surface Temperature Records

  • 1. Climatic Research Unit, School of Environmental Sciences, University of East Anglia, Norwich NR4 7TJ, UK
  • 2. Center of Excellence for Climate Change Research, Department of Meteorology, King Abdulaziz University, Jeddah, Saudi Arabia

Abstract: The purpose of this review article is to discuss the development and associated estimation of uncertainties in the global and hemispheric surface temperature records. The review begins by detailing the groups that produce surface temperature datasets. After discussing the reasons for similarities and differences between the various products, the main issues that must be addressed when deriving accurate estimates, particularly for hemispheric and global averages, are then considered. These issues are discussed in the order of their importance for temperature records at these spatial scales: biases in SST data, particularly before the 1940s; the exposure of land-based thermometers before the development of louvred screens in the late 19th century; and urbanization effects in some regions in recent decades. The homogeneity of land-based records is also discussed; however, at these large scales it is relatively unimportant. The article concludes by illustrating hemispheric and global temperature records from the four groups that produce series in near-real time.

1 Introduction
  • A number of groups routinely update gridded datasets of surface temperature for land and marine regions, which can be used to produce time series of global and hemispheric temperatures. The four main groups are: the UK Meteorological Office Hadley Centre/Climatic Research Unit, which produces HadCRUT4 (Morice2012; http://www.cru.uea.uk/cru/data/temperature/,and http://hadobs.metoffice.com/hadcrut4/)—an updated version of HadCRUT3 [Brohan et al., 2006]; the US National Centers for Environmental Information (NCEI; Karl 2015; https://www.ncdc.noaa.gov/climate-monitoring), which is an updated version of [Smith et al., 2008] and [Vose et al., 2012]; the Goddard Institute for Space Studies (GISS; Hansen 2010; http://data.giss.nasa.gov/gistemp/), which is an updated version of Hansen et al. (1999, 2006); and the Berkeley Earth Group (BEST; Rohde2013a, 2013b; http://berkeleyearth.org/). One other group monitors land-based temperatures [Lugina et al., 2006] and another monitors SST [Ishii et al., 2005]. Surface temperature datasets are comprised of measurements from the land (from air temperatures at fixed locations) and SST data from the ocean (from moving ships and buoys). SST data are used for the oceans instead of marine air temperatures (MAT, taken by ships), as a few SSTs in an area of ocean are much more reliable than the same number of MAT measurements. Additionally, the number of MAT measurements must be further reduced by half, due to daytime heating caused by the ship, so only night-time MAT (see discussion in Kent 2013) can be used. Data from the two components are combined as anomalies from a base period. The base period, however, is different for the four data sets: 1961-90 for HadCRUT4; 1901-2000 for NCEI; and 1951-80 for GISS and BEST.

    All the groups use much the same input data, but employ different approaches to interpolation to develop gridded products. HadCRUT4 and NCEI both use a 5º× 5 ºlatitudelongitude grid that is produced first for separate domains for land and ocean. These two gridded products have overlaps at coastlines and islands and are combined in different ways by the four groups. HadCRUT4 combines land data from CRUTEM4 [Jones et al., 2012] with SST data from HadSST3 (Kennedy2011a, 2011b; see also Kennedy2014). NCEI use land data from the ISTI database [Rennie et al., 2014] and their ERSSTv4 dataset of SST anomalies over the ocean (Huang2015; Liu2015). GISS data are derived by first averaging all the land station data (from NCEI) into 160 approximately equal-area boxes, and then combining these with SST values (currently using ERSSTv4) from marine areas. BEST uses numerous station datasets from NCEI, and also those used by CRUTEM4 combined with marine data from HadSST3.

    If there are no data for a given month in one of the grid boxes, the HadCRUT4 value is missing. All the other datasets perform some sort of spatial infilling to produce more globally complete fields-NCEI by using an eigenvector-based technique, where this is judged to produce statistically reliable estimates; GISS uses 160 equal-area boxes effectively to provide some infilling in data-sparse areas, so only a few of the boxes are completely missing for all months; and BEST use kriging procedures (see Rohde2013a, 2013b). The amount of infilling undertaken with NCEI, GISS and BEST is unknown without station coverage for each month/year. Maps of all stations used are unhelpful without knowing their data availability, especially before the 1950s. Additionally, a fifth group, the Japanese Meteorological Agency, combines the [Ishii et al., 2005] SST data with land stations from NCEI, http://ds.data.jma.go.jp/tcc/tcc/products/gwp/temp/ann_wld.html, but the method has not been formally published. Despite these differences in the methods used to combine the basic data, the hemispheric- and global-scale time series are very similar [see the trends calculated over three different periods in Table 3.3 of [Trenberth et al., 2007] for IPCC AR4, and Table 2.7 of [Hartmann et al., 2013] for IPCC AR5]. In section 7 of the present paper, trends for global averages over three periods (1901-2014, 1951-2014 and 1979-2014) are calculated.

    The purpose of this article is to first discuss (in section 2) the principal reasons for the similarities at large spatial scales, and then (in section 3) the important issues that need to be considered to ensure reliability and to assess the accuracy of the monthly and annual estimates (for hemispheric and global averages and also at the grid-box scale). Section 4 illustrates these for the biases (SSTs, exposure and urbanization), while land-station homogeneity is addressed in section 5. Section 6 discusses the results from a number of reanalyses of the climate system (e.g., ERA-Interim; Dee2011) in the context of changes in spatial coverage through time. With all this knowledge, section 7 discusses the hemispheric and global analyses produced by the four groups, and section 8 concludes.

2 Similarity and homogeneity of large area-average time series
  • There are three principal reasons for the close similarity between the four independently derived data series. The first is that they use much the same raw (monthly-mean) input data for the land areas and separately similar input data (International Comprehensive Ocean Atmosphere Dataset, ICOADS) for the marine areas. The second is that similar bias and homogeneity adjustments are applied to both sets of data, particularly for the ocean, and these form the main discussion points of this paper. While there are some minor differences in the input data and the adjustments applied, many of the homogeneity issues are essentially random; so, when averaged over large areas, the differences tend to cancel out. The third factor is often ignored and poorly understood; namely, that grid-box temperature time series from neighbouring locations are highly spatially correlated. Thus, even though there are records from thousands of sites on land and from millions of measurements from ships and buoys across the world’s oceans, the “effective” number is much less than this. Estimates (using both observational data and globally complete climate model data) indicate that the effective number of independent observations at the monthly timescale for the global surface area is about 100 (see Jones1997). Thus, provided input datasets have at least 100 well-spaced sampling points for which the data are relatively free of non-climatic biases, even if the locations of these sites are different between the different groups, they will lead to very similar large-scale area averages. For annual or decadal averages the required number of well-spaced locations can be substantially less than 100.

    A similar situation exists for pressure data. Here, the correlation decay length is similar to that for temperature, so relatively few sites can produce reliable area averages. For precipitation, however, the required number of data series to produce reliable area averages is much greater than for temperature, as correlation decay lengths are much smaller. The number of locations required to derive similar datasets from daily temperature averages would be larger, as at the daily timescale correlation decay lengths would also be smaller.

    The relatively small number of locations required to estimate large-scale area averages accurately means that, even for early parts of the temperature record when the data network was relatively sparse, area averages are reliable back to the second half of the 19th century. A test of the adequacy of the evolving network of temperature data sites for deriving large area-average time series is provided by Le Treut et al. (2007, Fig. 1.3). Here, many of the series developed before 1985 (all of which are just for the land regions of the world) are compared and shown to agree well. Even the record developed by [Koppen, 1873] for the Northern Hemisphere land masses is similar to averages developed today by CRUTEM4. The adequacy of the network used by Callendar (1938, 1961) is also excellent when compared to CRUTEM4 [Hawkins and Jones, 2013].

    The adequacy of early networks has also been illustrated using subsamples involving the use of present-day regions that had good sampling in the 19th century. [Parker et al., 2009], for example, have shown that the number of sites required to produce a reliable area average is small (see their Fig. 1). They did this by calculating global land averages using a limited set of 5 º grid boxes, and then with another analysis offset from the first by 10º of longitude and latitude. Earlier, [Jones, 1994] used a sparse but more constant network of stations to show that the sparser networks available in the second half of the 19th century could reproduce the global average reliably on decadal timescales and so ensure the consistency of large-scale area-average time series. Individual months and years may differ, particularly prior to 1900, but sparse networks are very reliable for decadal and longer-timescale averages.

    Network adequacy has also been discussed by [Cowtan and Way, 2014], who claim that HadCRUT4 underestimates warming in the last 15 years due to missing grid boxes in the Arctic. [Cowtan and Way, 2014] infill all missing grid boxes in HadCRUT4 from 1979 onwards using reanalysis products, lower tropospheric temperatures, or by kriging—not just across the Arctic, but also for the Antarctic, parts of Africa, and a few smaller areas. Using reanalysis can cause problems in the Antarctic [Jones and Lister, 2015] and is not to be recommended. Infilling by kriging tends to produce fields that are smoother than observed data show. In section 7, it is shown that the trends of the various datasets demonstrate that warming rates in HadCRUT4 are not significantly different from the other three datasets (see Table 1).

    The strong spatial correlation of temperature is also important in paleo-temperature reconstructions from proxy data. Here, the number of sites is much fewer than for instrumental data, but reliable area averages can still be produced (see Jansen 2007). Further back in time, millennial-and multi-millennial-scale temperature histories are derived from a few ice cores and/or deep sea cores (see Masson-Delmotte 2013).

3 Issues to consider in series adjustment and error assessment
  • The effective number of spatial degrees of freedom is one of the key parameters in estimating the statistical uncertainty in estimates of large-scale averages. In an earlier study by the HadCRUT group [Brohan et al., 2006], estimated uncertainties also account for uncertainties in homogeneity and bias adjustments to the basic data, possible urbanization influences, as well as the effects of sparser sampling in the earlier years. The incorporation of these components is complicated by the fact that some issues cancel by the number of measurements (particularly those due to land-station homogeneity), while the biases tend to be consistent so do not cancel. In order for uncertainty errors to be widely used, [Morice et al., 2012] introduced the concept of multiple, but equally plausible, realizations of the past. The HadCRUT4 dataset has developed 100 such realizations with a best guess, the median value for each grid box, and the median of the 100 realizations of global and hemispheric averages. The range of these realizations expands for years earlier in the record, but is still quite low in regions with good coverage back to 1850. [Smith and Reynolds, 2005] have also looked at the effects of sparser coverage in earlier years.

    Knowledge of the potential sources of error and their correlative structure is key information if the uncertainties in the global temperature record are to be reduced. The greatest potential for improvement will come from infilling data gaps in early years, particularly through the incorporation of more marine data [where improvements in metadata will also be important-as evidenced in the work of Thompson et al. (2008, 2009)]. As will be shown in this paper, the greatest uncertainty is in the marine data before World War 2, and this has recently been well illustrated by [Karl et al., 2015].

    To discuss the different uncertainty components, it is necessary to understand their structure; but before that, there is a need to define a few terms commonly used in climatology. There are three basic issues in the development of the gridded temperature products and global and hemispheric means: homogeneity of the basic raw station or marine time series; large-scale systematic biases that might affect large areas; and the lack of coverage in parts of the world, particularly before the 1950s. These will be discussed in the next sections in their order of importance for the large-scale averages: biases, coverage, and homogeneity of the individual site series. At the local (grid-box) scale, the order of importance would differ: coverage, then homogeneity, and finally bias. The fact that the order of importance depends on the spatial scale is a particularly vital aspect to realize. Additionally, the components of the uncertainty are independent of each other, so may be combined in quadrature (Brohan 2006; Morice2012), as opposed to being additive in nature.

    It is important to note, however, that these problems apply to the original (raw) input data. For the data that are used to produce standard area-average time series, corrections have been applied to remove, as far as possible, these potential sources of error. The fact that four different organizations have made such corrections independently is a testimony to the robustness and accuracy of the resulting homogenized data (see this illustrated in section 7). Related to this, adjustments for land data are estimated completely independently from the marine series, so these two components mutually support each other.

4 Biases
  • Biases are homogeneity issues that affect large portions of the observational dataset. They may be smaller in magnitude than the effects of site moves and other factors (see section 5), but they can be important if they similarly affect significant fractions of the basic input data. These will be discussed in order of importance, as measured by their impact on hemispheric-and global-average series. The three most important factors are: methods of measuring SST; exposure issues with temperature recorded at land stations before the development of louvred screens; and the time-varying effects of increasing urbanization due to the growth of cities (see also Jones 2010). This third factor is linked to the representativeness of the site in the context of possible land-use or environmental change across the grid box within which it is located. Land station homogeneity is discussed in section 5.

  • Any issue of homogeneity or bias in measuring SSTs will have a serious impact on global temperature estimates because almost 70% of the planet’s surface is ocean. The history of marine instrumental measurements goes back to the 18th century and the basic meteorological measurements (not just SST, but air temperature, pressure, wind direction and speed, etc.) were entered into ship logbooks. Even before instruments, ships kept logbooks as these were essential for navigation.

    The first SST measurements were taken using wooden buckets tied to a rope. A sample of sea water was hauled onto the ship’s deck and the temperature measured. In the earliest years these measurements came mostly from voyages of exploration. By the early 19th century a whole array of measurements, including SST became a routine part of life at sea [Maury, 1855]. The advent of steamships in the mid-19th century led to ships increasing their speed and deck height above the sea surface. By the late-19th century, many SST measurements were made with canvas buckets, which were more flexible and much cheaper to construct. The use of canvas buckets continued on most merchant and naval vessels up to about 1940. Bucket use continued after this, but designs were improved (see Kent 2010).

    When SST data were first examined in detail by climatologists in the 1970s (see references discussed by Folland 1995; Kent 2010; and Kennedy 2014), it was soon realized that the method of measurement might influence the results. Between the wars there were a number of comparisons made of different measurement techniques on research vessels and on cruises, i.e., comparisons of different types of bucket, as well as with thermometers fitted into the engine cooling-water intake pipes of ships (see Kent 2010; and Kennedy 2014 for details). More comparisons were undertaken in the 1960s and 1970s, and it was at this time that an extra code was added into both logbooks and transmitted data to indicate how the measurement was made [Woodruff et al., 2011].

    The different thermal properties of the buckets: wooden, canvas, and also, more recently, rubber, mean that to use these data it is necessary to determine their relative biases, and to develop a history of which types of bucket were used. Bucket-type biases have been extensively assessed by [Farmer et al., 1989] and [Folland and Parker, 1995]. Assessments are continuing as more of the history of recording is being found and more ship-logbook data digitized [see more recent discussion in [Ishii et al., 2005], Kennedy et al. (2011a, 2011b), [Kent et al., 2013], and [Kennedy, 2014].

    With regard to changing instrumentation, a basic assumption is that wooden buckets dominated in the 19th century, canvas buckets from 1900 to 1941, and engine intake measurements from then on. These were not, of course, abrupt changes, but spatially variable transitions over time, so correcting for these changes is not a simple task [Kennedy, 2014]. The importance of these measurement technique biases is evident from the average size of the adjustment across the world’s oceans-canvas bucket measurements need to be raised by about 0.4 ºC between 1900 and 1941 compared to engine intakes. The main cause here is the evaporative cooling of the sea water between the times of sampling and reading of the thermometer. The procedures provide corrections that can be applied for each part of the ocean with different values during the seasonal cycle. Temperatures measured in wooden buckets before the 1890s must also be raised relative to engine intake measurements, but by smaller amounts than for canvas buckets. Uncertainties in these adjustments are also incorporated in the overall error range accompanying each grid-box or larger-scale value [see discussion related to the multiple realizations in [Kennedy et al., 2011b] and [Morice et al., 2012]. These uncertainties are dependent on the size of the adjustments, so are larger for the canvas as opposed to wooden buckets. Thus, even though coverage is sparser in the late 19th century, the uncertainties are larger between 1910 than 1940 than those from the earlier sparser coverage.

    Although the major issues with SST data relate to the period before about 1940, there are still issues with recording in recent times. First, recent work has suggested that SST data for the period 1945 to 1960 are too cold (Thompson 2008, 2009). This is related to many of the measurements in this period being taken by British naval ships, which seem to have continued their canvas bucket method of sampling. These issues are being resolved with improved metadata and by attempting to relate individual measurements to the ships that took them (see discussion in Kennedy2014), but for about 30% of the SST observations during the period 1950-75, the measurement method is unknown. As stated earlier, buckets continued to be used after 1945, but designs were also much improved to minimize the bias (see Kent2010). Second, since the late 1970s, there have been major changes to the marine observing system, with the principal aim of improving weather forecasts and seasonal climate predictions. Satellites began to measure SSTs at this time, and fixed buoys have been deployed in the equatorial Pacific to help ENSO predictions. Further, since the late 1980s a large number of drifting buoys have been regularly deployed across the world’s oceans. Little consideration was given to the homogeneity of measurements at the time these instrumental changes and additions were made.

    As a consequence, when detailed comparisons have been made, potentially important inhomogeneities have been discovered [see discussion in [Kennedy, 2014], [Huang et al., 2015], and [Liu et al., 2015]]. For example, it seems that the new drifting buoys estimate SST values slightly lower than ships by 0.1ºC to 0.2 ºC, so their use might introduce a spurious cooling in the record. More extensive discussion of the SST adjustment procedures are provided by the different data centres [see Kennedy et al. (2011a, 2011b) and [Kennedy, 2014] for HadSST3, and [Huang et al., 2015] and [Liu et al., 2015] for ERSSTv4]. The recent study by [Karl et al., 2015] illustrates that SST adjustments are by far the largest factor impacting hemispheric and global temperature measurements. If the adjustments were not applied then century-timescale warming would be greater, and there would be a major discrepancy between the land and marine components prior to about 1940. This will be illustrated in section 7.

  • The problem of thermometer exposure, primarily to avoid the direct impact of sunshine on the instruments, was solved during the mid-to-late 19th century with the invention of screens. The problem had been recognized for many decades. Many different variants were tried, but it wasn’t until the development of white-louvred screens by Stevenson around 1870 (https://en.wikipedia.org/wiki/Stevenson_screen) that consistent exposures were established. Louvred screens have had different names around the world, e.g., “cotton region shelters” in the United States. Other changes to instrument exposure have also taken place in different regions around the world [see [Parker, 1994]; and [Trewin, 2010] for extensive discussions]. Prior to the development of screens, thermometers were generally positioned on north wall locations of buildings in the NH, so as not to be in direct sunlight. Despite this, they would have received some sun exposure during the early morning and late evenings during the summer, particularly the farther north the location. Before the use of screened locations, it was likely that air temperatures during the summer half of the year could be biased warm. Measurements during the winter half of the year would be unaffected.

    Although the long-term homogeneity of station temperature series can be assessed (see section 5 for more discussion of this issue), the accuracy of these approaches in early years in Europe has been questioned by some authors, particularly for measurements made during the summer (see Moberg 2003; Jones2003), as all series are similarly affected by screens being introduced in some regions at the same time. The discussion in these papers has centred on two aspects of long temperature series: (1) the warmth of summer temperatures in the pre-1880 period; and (2) the lower long-term warming rates in summer compared to the other three seasons. Crucially, if these issues are important, it will be regional in nature (especially in Europe and for the early parts of individual station records), but they will be of little significance for global-scale changes over the period since 1880.

    [Moberg et al., 2003] showed that the long-term warming in Swedish winters is consistent with changes in the atmospheric circulation (the North Atlantic Oscillation in this case) and the warming of SSTs in the North Atlantic, so is likely to be reliable. The reliability of the summer data is harder to determine, however. The circulation and local SSTs influence summer temperatures considerably less, and the principal determining factor here is the radiation received. This can be estimated from long series of cloudiness data but the long-term homogeneity of cloudiness data from early observers is beset with even more problems than for temperatures, and so cannot help to assess the reliability of long summer temperature series.

    When any change to observing practice takes place, it is always recommended that parallel measurements are made [see the GCOS Monitoring Principles in [Bojinski et al., 2014]. This doesn’t always happen, and even if it did in the 19th century, the comparison measurements have probably not survived. Climatologists have recently begun to collect modern parallel measurements to attempt to resolve these instrument exposure issues. Two examples of this type of work are studies in the Greater Alpine Region (GAR) by [Bohm et al., 2010] and in Spain by [Brunet et al., 2011]. The former used parallel measurements at one site in Austria, which enabled the differences between the old and modern exposure methods to be estimated and related to the directional exposure of all earlier sites in the GAR. The Spanish example rebuilt screens from 19th century diagrams and made modern parallel measurements, again developing correction formulae to apply to the original 19th century data. The results were similar in both cases—summer temperatures on average were recorded about 0.4ºC warmer with the old, compared to the modern, exposures. These results are very similar to pioneering assessments made at Adelaide in Australia [Nicholls et al., 1996]. It is believed that instrumental series across much of Australia before 1910 are affected [Trewin, 2010].

    Assessment of early instrumental exposure is vital not just for long-term homogeneity, but also for the response of many natural proxies (particularly trees) to summer temperatures. Temperature reconstructions from proxy data records clearly require the primary temperature data against which the proxy data are calibrated to be reliable. If the homogeneity of pre-1900 temperature records from individual sites could be improved, this could enhance the reliability of such temperature reconstructions.

  • Station history information often shows that many sites began at locations in small towns, and that, over the last 100 years, some of these towns have developed into major cities. Such urban growth is likely to affect temperature records from urban sites, and warming trends from such sites are likely, on average, to be larger than if the city or town were not there (see review by Arnfield 2003). In climatology, this issue is referred to as the urbanization effect or the urban heat island. The implication of this effect for gridded datasets is that urban-affected sites will no longer be representative of the majority of the grid box. This could potentially impact large-scale temperature averages if gradually more of the sites during the 20th century are located in urban areas. The issue is not the urbanization effect per se, but whether nearby rural and urban locations show similar long-term trends. For example, city centre sites in London and Vienna are warmer than their rural counterparts, but the urban time series during the 20th century change at exactly the same rate (see Jones 2008; Jones 2009).

    Numerous papers have addressed urban climates and found large differences between city centre sites and rural neighbours for individual day and night temperatures (see references in Arnfield 2003) but these studies are generally not relevant to the global-scale data bases described here. This is because most of these comparisons only consider days that maximise the urban/rural difference and so are not directly relevant in the context of long-term monthly averages for typical (non-city-centre) weather stations. Using the example from London [Wilby et al., 2011] an urbanization effect over decadal timescales is apparent, but this could easily be explained by some periods being dominated by circulation patterns that emphasize an effect while other patterns reduce the effect.

    There are a number of other factors that make the assessment of urbanization effects difficult, but as shown below, it is likely that residual errors are small. The first factor that must be considered is that sites in urban areas are generally not in the downtown part of the city, but are more likely to be in a parkland setting or at an airport (see Peterson 2005). The second is that each case probably has to be assessed individually. It is impossible to make generalizations: European cities, for example, will differ from cities in other parts of the world.

    Despite the difficulties in correcting for urbanization effects, there are two strong arguments that indicate that any residual urbanization effects in the standard (homogenized) temperature datasets are probably very small. The first of these is that SSTs are not affected, so the similarity of warming trends from land and marine regions argues against the effect being important. Second, datasets can be constructed using only rural locations. Although this restricts coverage, because of the spatial correlations, sparser networks can be used to derive reliable large-scale averages [see section 2 about earlier discussion on spatial degrees of freedom in [Jones et al., 1997]. When compared with results using many stations, the differences are small [see the review by [Parker, 2010]].

    As noted above, many assessments of urbanization effects at the large scale have considered rural-only sites and compared these to averages based on all sites, or on urban-only sites [see, for example, [Jones et al., 1990], Parker (2004, 2006, 2010), and [Peterson and Owen, 2005]]. Differences are always small, and always an order of magnitude smaller than any long-term warming—implying that any urbanization effect is small. [Wickham et al., 2013] assessed all the land stations in the BEST land station dataset, putting each station into one of two groups (“very rural” and “not very rural”) depending on the land-use around the station location. Interpolating these two categories separately, they found no statistically significant differences in their two global average series.

    Locally, however, the effects may be larger, and much recent work has emphasized China (e.g., Ren 2008) because here the effect may be larger than in other parts of the world [Jones et al., 2008]. Urban growth has been dramatic in the recent 30 years across eastern China, but an important consideration is that very few of the series are located in rural locations [see discussion in recent papers by [Li et al., 2014] and [Zhao et al., 2014]]. As in other parts of the world, the issue that is especially important in China is the representativeness of the network, particularly for locations that are distant from the measuring sites. Do urban sites represent rural regions in eastern China? Averages produced for China or parts of China, e.g., by [Li et al., 2014] and [Zhao et al., 2014], use networks of different station densities, including both rural and urban stations. Others (e.g., Ren 2008) omit the more urban stations. For both types, averaging doesn’t consider land use except at the stations. [Wang et al., 2015] addressed this issue in a different way in a recent study by considering land-use information across China for the period since 1980 and determined an urban land index for each of their 607 stations across the country. Stations were then divided into three categories (intense, moderate and minimal urbanization) and each of the three groups was used separately in developing gridded products (for a 2.5º ×2.5º latitude-longitude grid). A China average was then calculated according to the proportion of urban land index across the country. The simple average of all the stations shows a greater warming than the land-use weighted series because urban areas (which constitute less than 1% of the total area of the country) are where 68% of stations are located. In summary, there is an urbanization effect in eastern China, but its impact could be considerably reduced by using a network of rural sites.

5 Homogeneity of individual land-based records
  • Individual temperature records from land sites are homogeneous [Conrad and Pollak, 1962] if the variations in the measurements result solely from regional-scale variations (at the scale of 10º×10º of latitude-longitude) in the weather and climate. Inhomogeneities result from many factors, some of which (instrument exposure, urbanization) have already been discussed. In addition, individual records may be affected by changes in site location, changes in the times each day the measurements are made, changes in the method used to calculate daily-and hence monthly-mean temperatures, and changes in instrumentation [see the recent review by [Trewin, 2010]].

    Several homogenization algorithms have been identified and assessed in recent years [see [Venema et al., 2012] for comparisons of the methods]. Once inhomogeneities are identified, the raw individual site records need to be adjusted to produce homogeneous time series. Adjustment factors are determined using station histories and metadata information (where this is available). Both physically-based corrections and corrections derived from objective statistical tests (comparing temperature time series from neighbouring sites) are estimated. Where necessary, adjustment factors are then calculated and the early parts of the records are made compatible with the most recent data. Additionally, those methods also calculate the uncertainties of the adjustments. While the effects of inhomogeneities vary from site to site, occasionally all the sites within a particular country may be affected (changes to exposure and urbanization both fall into this category in some countries). When this happens, homogeneity assessment using neighbours may not work well, as all series are likely to be similarly affected.

    For individual site records and for small-scale averages (such as at the single grid box level), homogenization is essential. As stated earlier, at this scale, site homogeneity issues are likely the most important of all the factors. At the hemispheric and global scale, however, because adjustments of both signs occur with similar frequencies, the adjustment factors tend to cancel. While there are uncertainties in adjustments at the site level, at larger scales the effects of such uncertainties are small compared to the SST biases and exposure issues (see section 4). The cancelling can be easily seen in a number of recent papers [e.g., Brohan et al. (2006, Fig. 4), Menne et al. (2009, Fig. 6), and locally for China in Xu et al. (2013, Fig. 3)]. Each of these figures shows histogram counts of the magnitude of adjustments, with the first two showing bimodal distributions with peaks for both positive and negative adjustments. The overall average of adjustments across multiple sites in a region is essentially zero. As station homogeneity is important at the local scale, adjustments are still made for individual sites since these are necessary to produce the best-possible gridded data.

    Figure 1.  Hemispheric and global averages, based on land and marine data, from the four datasets discussed in this paper: HadCRUT4 [Morice et al., 2012]; NCEI/NOAA [Karl et al., 2015]; GISS [Hansen et al., 2010]; and Berkeley Earth (http://berkeleyearth.lbl.gov/auto/Global/Land_and_Ocean_summary.txt). The HadCRUT4 range encompasses the 5% and 95% values from their 100 ensembles [Morice et al., 2012]. The unadjusted data are from NCEI [Karl et al., 2015]. All data are expressed as anomalies from the 1961-90 average.

    A more recent example of changes in instrumentation is the automation of measurements across whole countries and regions that has taken place during the last 25 years [e.g., for the U.S., in [Quayle et al., 1991]]. It is, however, possible to identify such changes and correct for them, provided dates of the changes are known. Another example from the USA is the change in observation time of daily maximum and minimum temperatures from late afternoon to early morning, which has been referred to as “time of observation bias” (TOB) and corrected for by [Karl et al., 1986]. The effect is noticeable because morning readings tend to be slightly cooler than those taken in the late afternoon. Figure 4 in [Menne et al., 2009] shows the effect of the TOB for the contiguous U.S. average from 1900, amounting to a difference of about 0.2{Invalid MML}C between adjusted and unadjusted data during the present decade. In other words, the TOB leads to a spurious cooling trend in the unadjusted data.

6 Comparison with reanalyses
  • Atmospheric reanalyses have been produced since the mid-1990s [Kalnay et al., 1996], and these potentially provide a means to assess gridded products of surface temperature. The most comprehensive current reanalysis (ERA-Interim; Dee 2011) is in excellent agreement with surface temperature datasets (see Simmons 2010), but this is not unexpected as this reanalysis assimilates surface temperature data. Extended reanalyses, e.g., 20CR [Twentieth Century Reanalysis[Compo et al., 2011]] and ERA-20C [Poli et al., 2013], only assimilate surface pressure data. They have been given similar SST data for the world’s oceans, so comparisons with gridded surface temperature products need to be restricted to the terrestrial regions. Agreement is excellent [see, for example, Compo et al. (2011, 2013) and [Parker, 2011] for 20CR, and also [Poli et al., 2013] for ERA-20C and [Hersbach et al., 2015] for ERA-20CM], which attests to the reliability of both the terrestrial surface air temperature data and the driving SST data. If the latter had not been adjusted for the large bias due to the change from bucket measurements, then the agreement with the land record would not have been produced. [Folland, 2005] illustrated this by forcing an atmospheric GCM with adjusted and unadjusted SST data from HadSST2 [Brohan et al., 2006]. Air temperatures over land areas forced by unadjusted SSTs were incompatible with observed air temperatures over land areas. Differences were clearest in the less variable regions of the world, such as the tropics.

    Reanalysis products have also been used to assess potential urbanization effects in surface air temperatures over land areas, particularly over China. The assumption here is that reanalyses do not know about changes in land use. Initial work in this area was suggestive of a large effect [e.g., [Zhou et al., 2004] for southern China], but more detailed studies over different parts of China and for different periods [Wang et al., 2013] showed results were very susceptible to the choices of region and period.

7 Comparison of hemispheric and global averages
  • Figure 1 shows hemispheric and global averages from the four groups, with results expressed as anomalies from the 1961-90 base period used by HadCRUT4. The uncertainty estimates from HadCRUT4 show the 5th and 95th percentile range based on the 100 ensembles of the uncertainty components [Morice et al., 2012]. For the NCEI/NOAA analysis, the additional analysis using unadjusted data (for both the land and marine components) is also shown (from Karl 2015). The Berkeley Earth analysis only produces a global average (http://berkeleyearth.lbl.gov/auto/Global/Land_and_Ocean_summary.txt). For the NH, agreement is excellent with the NCEI/NOAA and GISS series within the HadCRUT4 uncertainty range. This uncertainty range expands before 1950 as slightly more of the NH has missing coverage. For the SH, error ranges for HadCRUT4 are wider than for the NH, reflecting the greater area of missing data coverage for HadCRUT4. Both NCEI/NOAA and GISS, for the SH, are near the lower uncertainty range (5th percentile) for the period from about 1920 to 1940 and from 1945 to 1965. As both these datasets use ERSSTv4, this is a result of different adjustment procedures for SST compared to HadSST3. HadSST3 assumed more of the SST measurements during these periods were from canvas buckets, particularly the latter period (see Kennedy 2011b; Thompson 2008, 2009). Getting SSTs correct in the SH is more important there than for the NH.

    Figure 2.  Hemispheric and global averages, based on land data, from the four datasets discussed in this paper: CRUTEM4 [Jones et al., 2012]; NCEI/NOAA [Karl et al., 2015]; GISS [Hansen et al., 2010]; and Berkeley Earth (http://berkeleyearth.lbl.gov/regions/global-land). The CRUTEM4 range encompasses the 5% and 95% values from their 100 ensembles [Morice et al., 2012]. The unadjusted data are from NCEI [Karl et al., 2015]. All data are expressed as anomalies from the 1961-90 average.

    Figure 3.  Hemispheric and global averages, based on marine data, from two of the datasets discussed in this paper: HadSST3 [Kennedy et al., 2011b] and NCEI/NOAA [Karl et al., 2015]. The HadSST3 range encompasses the 5% and 95% values from their 100 ensembles [Morice et al., 2012]. The unadjusted data are from NCEI [Karl et al., 2015]. All data are expressed as anomalies from the 1961-90 average.

    The global average is (NH + SH)/2, but the greater interannual variability of the NH tends to dominate. BEST is only available for the global average. The BEST series follows HadCRUT4, principally due to their common use of HadSST3 for ocean areas. Despite this, BEST implies cooler temperatures during the period before about 1890—a feature which must be related to cooler land temperature anomalies than HadCRUT4. Finally, the unadjusted NCEI/NOAA data imply much cooler temperatures before 1940, as canvas bucket adjustments were not applied (see also Karl2015). To further illustrate the importance of the ocean adjustments, Figs. 2 and 3 are similar to Fig. 1 but show hemispheric and global plots for the land (Fig. 2) and marine (Fig. 3) parts of the world. The unadjusted NCEI/NOAA data for the land areas of the world (Fig. 2) are not distinguishable from the CRUTEM4, NCEI/NOAA (adjusted), GISS and BEST time series. Minor differences occur, but they are within the CRUTEM4 5%/95% uncertainty ranges. For marine regions (Fig. 3), the unadjusted NCEI/NOAA marine data are clearly offset (for periods before the 1960s) from their adjusted data (ERSSTv4) and HadSST3, and fall outside the 5%/95% uncertainty ranges based on HadSST3. Furthermore, the difference between ERSSTv4 and HadSST3 is quite large at times, particularly for the SH (e.g., for the 1930s and the 1950s)—clear evidence that the uncertainty in SST bias adjustment is much larger than for the terrestrial part of the world in Fig. 2.

    On interannual timescales in all three figures, warm years can be clearly related to El Niño years and cool years to La Niña years or to large explosive volcanic eruptions in the tropics [see illustrations of this in [Foster and Rahmstorf, 2011]]. The greatest El Niño events of the last 200 years occurred in 1877/78 and 1997/98. On longer timescales, the world has warmed in two phases, from about 1920 to the early 1940s and from the late-1970s. The warmest year in all four global records is 2014, but this value only just exceeds that measured in 1998, 2005 and 2010. Initial data for 2015, partly due to the current El Niño, indicate that 2015 will be significantly warmer than all other years. If the El Niño event continues, then it is possible that 2016 will be warmer still.

    Finally, in this section, trends are calculated for the global average (from the land and marine datasets) for the four datasets and for NCEI-unadjusted for three different time periods (1901-2014, 1951-2014 and 1979-2014). The final period represents the period of satellite coverage. The results are given in Table 1 and all trends are statistically significant at the 99% level for all periods. The NCEI uncorrected series is also included and this clearly shows a greater long-term warming (for 1901-2014) than would have occurred if the bias and homogeneity adjustments were not applied. The results presented here and in [Karl et al., 2015] and [Kennedy et al., 2011b] clearly show that this is due to the SST bias (see Fig. 3).

    Much has been written about temperature trends over the past 15 years (often starting during or just after the major El Niño event of 1997/98), with the period being referred to as a “ hiatus” in warming (e.g., Hartmann2013; Karl2015, and references therein). A number of possible explanations have proposed for this, but [Karl et al., 2015] conclude that their new analysis doesn’t support the notion of a hiatus. From a data perspective, this will be further enhanced by the upcoming warm years of 2015 and 2016. As La Niña events generally follow El Niño events, it is likely that 2017 and 2018 might be cooler. Rather than then starting a new hiatus, it could be beneficial to additionally discuss global average temperatures after the effects of El Niño and La Niña events have been removed [using approaches similar to [Thompson et al., 2009] or [Foster and Rahmstorf, 2011]].

8 Conclusions
  • The importance of inhomogeneities in raw surface temperature observations becomes clear when comprehensive models to estimate the uncertainties involved are developed (e.g., Brohan 2006; Morice2012; Karl 2015). Factors that affect individual site records tend to be random (i.e., they can lead to positive or negative biases) and so uncertainties in any adjustments for land stations become less and less important as data are averaged over larger areas. Biases that affect multiple sites or records (such as changing measurement techniques for SSTs, changes in exposure of land stations and urbanization), although smaller in magnitude than many individual land station adjustments, become more important the larger the area averaged. As illustrated by Fig. 1, the four groups independently account for all these issues and produce series within the error estimates of HadCRUT4. Using only unadjusted data, [Karl et al., 2015] show that if the biases and homogeneity issues are ignored, the world would have warmed more. This result is primarily due to the SST bucket bias.

    The impacts of sparser coverage in early decades are only important before 1880, and, even then, the impact is mostly felt in the Southern Hemisphere [Jones, 1994]. For the Northern Hemisphere, it is possible to derive reliable hemispheric averages from instrumental data back to about 1850. For example, [Karl et al., 1994] show that global 100+ year trends become quite reliable after the 1870s based on historical sampling.

    Understanding the major sources of inhomogeneity provides key information for reducing uncertainties in hemispheric averages. Uncertainties would be most significantly reduced through the inclusion of more SST data in the 19th century than through adding more land station series since the 1950s. A number of current projects are seeking to digitize much of the British logbook material available in archives. The potential size and importance of SST data, also requires enhancements to our knowledge of how SST and MAT measurements were taken in the past (Kent 2010; Kennedy 2014). More SST data are not only important for improving the reliability of hemispheric and global temperature series, but can help to improve infilled SST fields, which are vital for extended reanalyses. For terrestrial regions, adding more land stations can also help reduce uncertainties, but emphasis needs to be focussed on regions with sparse coverage, as opposed to simply increasing station numbers in well-monitored regions. For identifying past large-scale changes in temperature at the Earth’s surface, however, the homogenized datasets currently available provide highly reliable information back into the 19th century and show unequivocally that the world has warmed considerably over this period.

    Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return