Elsevier

Atmospheric Environment

Volume 37, Issue 23, July 2003, Pages 3167-3175
Atmospheric Environment

Long-term variations in total ozone derived from Dobson and satellite data

https://doi.org/10.1016/S1352-2310(03)00347-9Get rights and content

Abstract

Total ozone growth rates are calculated using flexible ‘tendency curves’ that can follow ozone variations on all timescales greater than that of the quasi-biennial oscillation. This method improves on traditional trend analysis using straight line fits because it follows ozone variations more closely, providing visual information about the timing and global distribution of ozone variations. Results are compared from long-running Dobson sites and from two homogenized satellite data sets, one constructed at NASA/Goddard Space Flight Center and the other developed at New Zealand's National Institute of Water and Atmospheric Research. Although the most negative ozone trends in the Southern Hemisphere appear to be linked to polar vortex chemistry, those in the Northern Hemisphere have occurred between 35°N and 40°N and may be related to dynamical trends and/or chemistry on episodically occurring volcanic aerosols. A quasi-decadal cycle in total ozone was present since the mid-1920s and hence is independent of halogen chemistry. Its cause remains unknown. Including the deseasonalized and detrended local temperature in the ozone trend model decreases the standard error of the ozone trend over most of the globe.

Introduction

Variations in total ozone have received much scientific scrutiny since the 1970s when it was first proposed that man-made chemicals such as chlorofluorocarbons (CFC) could impact the Earth's protective ozone shield (Molina and Rowland, 1974). Decreases in the stratospheric ozone column allow more ultraviolet radiation in the 280–320 nm range to reach the surface where it is harmful to living organisms. Trend studies gained prominence when it was recognized that an ozone ‘hole’ was forming every spring over Antarctica (Farman et al., 1985). The word trend has implied a ‘generally consistent, continuing change over a given period’ (WMO (World Meteorological Organization), 1995). Bojkov et al. (1996) referred to “a process of ozone decline that has a stable and continuous cause”. Some past trend studies have looked for a linear response beginning in 1970 to increases in CFC emissions such as one might expect from gas phase halogen chemistry. However, calculated ozone trends were much larger than the 1% change expected from homogeneous chemistry above 30 km. Furthermore, the majority of ozone loss was in the lower stratosphere below 25 km (Stolarski et al., 1992). Past trend studies also found that the interval chosen for trend determination often had a significant impact on trend values, so that total ozone variations with time were in fact not linear. Solomon et al. (1996) identified large changes in stratospheric aerosol abundance in the presence of halogens as a source for some of those non-linear ozone losses. In the latter study the magnitude of modeled changes was 50% smaller than those observed in total ozone, but the timing and location of the losses was correct. Thus, the accelerated losses in total ozone in the Northern Hemisphere midlatitudes seen after the Pinatubo volcanic eruption may have been caused by heterogeneous chlorine and bromine chemistry occurring on liquid sulfate particles in the lower stratosphere.

Another source for ozone depletion that does not follow a linear model is dilution from the Polar Regions. This process occurs when filaments of ozone depleted air are transported into midlatitudes when the vortex erodes in the spring. It is especially evident in the Southern Hemisphere, where it may have a cumulative effect from one year to the next which is estimated to add about 20% to the depletion (Prather et al., 1990).

The Montreal Protocol and its amendments have led to a reduction in halogen emissions so that concentrations in the stratosphere are no longer increasing in a linear fashion. Aside from halogen chemistry, other effects that one would not expect to be linear are those driven by changes in atmospheric dynamics. Recent studies provide evidence for the role of dynamics in long-term midlatitude total ozone variations (Reid et al., 2000; Thompson et al., 2000; Appenzeller et al., 2000). Increases in greenhouse gases and water vapor may contribute to depletion by causing radiative cooling of the stratosphere (Shindell et al., 1998; Dvortsov and Solomon, 2001).

The unfolding knowledge of different sources for midlatitude total ozone variations bring into question the assumption that long-term total ozone variations should be modeled with a straight line. Fioletov et al. (2002) has dealt with the issue of nonlinear ozone variations by calculating sliding 11-year trends. Their method successfully shows how trends change with time, although it still assumes linear behavior over each 11-year period. In this study a flexible curve is fitted to ozone residuals after known natural variations are removed. This curve, referred to here as the total ozone tendency curve, can follow decreases, increases, and accelerations so that it gives more visual information than a straight line. The derivative of this flexible curve is the growth rate curve, whose average value is comparable to the trend values of past studies. Although we no longer assume that the ozone trends solely reflect the effect of homogeneous halogen chemistry on ozone, calculating these trends is still an important task and allows comparison with other studies. The flexible tendency and growth rate curves give specific information in terms of the timing and location of total ozone variations over the globe. Applied to satellite data, this method illuminates geographical differences in ozone variation, thereby giving clues to possible processes affecting total ozone. In the future these visual tools may be used to monitor for unexpected changes, eventual recovery, and possible effects of climate change.

Section snippets

Total ozone data sets and trend analysis method

This paper presents total ozone growth rates derived from three sources: Dobson monthly data and two homogenized satellite monthly data sets. Dobson column ozone measurements with long-term (30+ years) records from the NOAA/CMDL Cooperative Dobson Network are analyzed. In addition, the record for Arosa, Switzerland, going back to 1926 is employed in this study. Representing the Southern Hemisphere are four sites with records that began prior to the satellite data. Melbourne and Lauder records

Total ozone growth rates—site comparisons

Table 1 lists the average growth rates for the Dobson sites, calculated for July 1967 through December 2001 and for November 1978 through December 2001. In addition, growth rates are reported for the GSFC and NIWA data sets for the grid boxes closest to the Dobson sites for the purposes of comparison. The Dobson sites have average growth rates from −1 to −2%/decade for the longer period except for Buenos Aires whose growth rate is not statistically significant. Considering the trends from

Conclusions

Trends in total ozone from November 1978 through 2001 have been determined from three sources: Dobson data and two homogenized satellite data sets, one calibrated to EP-TOMS (GSFC) and the other scaled to Dobson measurements (NIWA). The greatest discrepancy in trends calculated from these three sources was on the order of 1%/decade. The NIWA growth rates are uniformly more negative than the GSFC growth rates.

Northern midlatitude trends calculated from 1978 are generally from −2 to −3%/decade.

References (23)

  • G. Reinsel et al.

    Statistical analysis of stratospheric ozone data for the detection of trends

    Atmospheric Environment

    (1981)
  • C. Appenzeller et al.

    North Atlantic oscillation modulates total ozone winter trends

    Geophysical Research Letters

    (2000)
  • G.E. Bodeker et al.

    Global ozone trends in potential vorticity corrdinates using TOMS and GOME intercompared against the Dobson network1978–1998

    Journal of Geophysical Research

    (2001)
  • G.E. Bodeker et al.

    The global mass of ozone1978–1998

    Geophysical Research Letters

    (2001)
  • R.D. Bojkov et al.

    Total ozone trends from quality-controlled ground-based data (1964–1994)

    Journal of Geophysical Research

    (1996)
  • V.L. Dvortsov et al.

    Response of the stratospheric temperatures and ozone to past and future increases in stratospheric humidity

    Journal of Geophysical Research

    (2001)
  • J.C. Farman et al.

    Large losses of total ozone in Antarctica reveal seasonal ClOx/NOx interaction

    Nature

    (1985)
  • Fioletov, V.E., Bodeker, G.E., Miller, A.J., McPeters, R.D., Stolarski, R., 2002. Global and zonal ozone variations...
  • J.M. Harris et al.

    A new method for describing long-term changes in total ozone

    Geophysical Research Letters

    (2001)
  • E. Kalnay

    The NCEP/NCAR reanalysis 40-year project

    Bulletin of the American Meteorological Society

    (1996)
  • J. Lean

    Evolution of the sun's spectral irradiance since the maunder minimum

    Geophysical Research Letters

    (2000)
  • Cited by (18)

    • Trend of total column ozone over Mexico from TOMS and OMI data (1978-2013)

      2014, Atmosfera
      Citation Excerpt :

      For this work at the OMI phase, only measurements with a step of 1° longitude × 1° latitude were used. TOMS and OMI satellite measurements have been validated against ground-based measurements and have become a standard long-record reference in many studies (McPeters and Labow, 1996; Fioletov et al., 1999; Bodeker et al., 2001; Harris et al., 2003; Jaross et al., 2003; McPeters et al., 2008). In order to take advantage of TOMS and OMI measurements, the understanding of the structure of the digital files is required, as well as facilities in order to read the data of interest.

    • Trends in total ozone column over India: 1979-2008

      2011, Atmospheric Environment
      Citation Excerpt :

      Initial attempts in analyzing the long-term trend in the TOC contents revealed an increase in its abundance over many locations (Komhyr et al., 1971; Angell and Korshover, 1973; London and Kelley, 1974). However, investigations undertaken after the discovery of “Ozone Hole” in 1985 (Farman et al., 1985), involving statistical analysis of TOC data to estimate the long-term trends in the TOC reported a decline: data used in these analysis was post 1970s onwards for many locations (Reinsel et al., 1987, 1994; Bojkov et al., 1990, 1995; Bojkov and Fioletov, 1995; Harris et al., 2003). The analysis included representative data over locations in different climatic zones, but the stress was more on the temperate and Polar Regions where the extent of variability in TOC abundance is large.

    • Ozone recovery as seen in perspective of the Dobson spectrophotometer measurements at Belsk (52°N, 21°E) in the period 1963-2008

      2009, Atmospheric Environment
      Citation Excerpt :

      This curve is therefore interpreted as the anthropogenic component of the ozone variability. The trend model having the trend pattern not a priori defined, called Flexible model further in the text, was introduced by Harris et al. (2001, 2003). The flexible trend model including MARS technique has been widely examined in our previous papers (Krzyścin et al., 2005; Krzyścin, 2006; Krzyścin and Borkowski, 2008).

    • Chemical components from the surface of Haplopappus bustillosianus

      2007, Biochemical Systematics and Ecology
      Citation Excerpt :

      The extremely low yield of n-alkanes in H. bustillosianus is, then, consistent with its environmental conditions since this species is not susceptible to suffering water stress. An average ozone decline of around 2.5% per decade has been estimated for the band from 30° to 50° in the southern hemisphere (Harris et al., 2003), where Lake Viillarica is located (39°16′ S). Also, recurrent episodes of ozone-poor air masses reaching mid-latitudes in South America have been documented (Lovengreen et al., 2000; Perez and Jaque, 1998).

    View all citing articles on Scopus
    View full text