Long-term (2004–2015) tendencies and variabilities of tropical UTLS water vapor mixing ratio and temperature observed by AURA/MLS using multivariate regression analysis

https://doi.org/10.1016/j.jastp.2016.08.001Get rights and content

Highlights

  • Increasing (decreasing) trend in WVMR (temperature) above 82 hPa (56 hPa).

  • No significant trend in WVMR and temperature near tropopause (100 hPa).

  • Low tropopause temperature and stratospheric WVMR during 50 hPa westward QBO.

  • During El Niño, negative anomaly in tropopause temperature throughout Pacific.

  • This explains negative response of stratospheric WVMR to El Niño.

Abstract

Long-term variabilities and tendencies in the tropical (30°N–30°S)monthly averaged zonal mean water vapor mixing ratio (WVMR) and temperature in the upper troposphere and lower stratosphere (UTLS), obtained from the Microwave Limb Sounder (MLS) instrument onboard Earth Observing System (EOS) satellite for the period October 2004-September 2015, are studied using multivariate regression analysis. It is found that the WVMR shows a decreasing trend of 0.02–0.1 ppmv/year in WVMR below 100 hPa while the trend is positive (0.02–0.035 ppmv/year) above 100 hPa. There is no significant trend at 121 hPa. The WVMR response to solar cycle (SC) is negative below 21 hPa. However, the magnitude decreases with height from 0.13 ppmv/100 sfu(solar flux unit) at 178 hPa to 0.07 ppmv/100sfuat 26 hPa. The response of WVMR to multivariate El Niño index (MEI), which is a proxy for El Niño Southern Oscillation (ENSO), is positive at and below 100 hPa and negative above 100 hPa. It is negative at 56–46 hPa with maximum value of 0.1 ppmv/MEI at 56 hPa. Large positive (negative) quasi-biennial oscillation (QBO) in WVMR at 56–68 hPa reconstructed from the regression analysis coincide with eastward (westward) to westward (eastward) transition of QBO winds at that level. The trend in zonal mean tropical temperature is negative above 56 hPa with magnitude increasing with height. The maximum negative trend of 0.05 K/year is observed at 21–17 hPa and the trend insignificant around tropopause. The response of temperature to SC is negative in the UTLS region and to ENSO is positive below 100 hPa and mostly negative above 100 hPa. The negative response of WVMR to MEI in the stratosphere is suggested to be due to the extended cold trap of tropopause temperature during El Niño years that might have controlled the water vapor entry into the stratosphere. The WVMR response to residual vertical velocity at 70 hPa is positive in the stratosphere, whereas the temperature response is positive in the UTLS region and negative above 56 hPa. Besides, the interannual variability and the response of the WVMR to the different parameters are explained based on the response of temperature at 100 hPa (proxy for tropopause) to those parameters.

Introduction

Though the water vapor content in stratosphere is small, it plays a major role in determining the radiative energy balance (Forster and Shine, 1999, Foster et al., 2002), in acting as a source of hydroxyl radicals and in the destruction of polar ozone by involving in the activation of chlorine on polar stratospheric clouds (Solomon, 1999). Water vapor is important in heterogeneous ozone loss in the polar regions because it lowers the threshold for formation of polar stratospheric clouds. More important perhaps, it also increases the heterogeneous reactivity of key reactions in polar ozone loss (Drdla and Muller, 2012). Brewer (1949) suggested that the water vapor in the lower stratosphere must have passed through the cold tropopause region over the tropics, slowly ascending within the circulation that later became known as Brewer–Dobson circulation (BDC). Methane oxidation is an important source of water vapor in the middle stratosphere (Jones and Pyle, 1984, Rohs et al., 2006). The tropical tropopause temperature controls the water vapor entering into the stratosphere (Fueglistaler et al., 2009). The annual variation in stratospheric water vapor is a response to the annual cycle in tropopause temperatures (Mote et al., 1996, Holton et al., 1995). Increase of tropospheric temperature leads to high stratospheric water vapor and through feedback process to further warming of troposphere (Dessler et al., 2013). The large positive anomaly of stratospheric WVMR during 1997–98 was attributed to the El-Niño event, which warmed the tropical tropospheric temperature by 2 K (Randel et al., 2004). The interannual variability of stratospheric water vapor is less, when compared to its annual changes and is mainly governed by quasi-biennial oscillation (QBO) (e.g. Giorgetta and Bengtsson, 1999), El-Niño Southern Oscillation (ENSO), and Brewer–Dobson circulation (BDC) (Randel et al., 2006, Dhomse et al., 2008). Thermal-wind relationship of QBO suggests warmer tropopause temperature during the eastward shear of QBO winds leading to more stratospheric water vapor, when compared to the westward shear of QBO (Baldwin et al., 2001). During winter, water vapor along with other chemical constituents get transported from the upper troposphere to extra-tropical stratosphere by BDC, which is forced mainly by the breaking of planetary-scale Rossby waves mostly at mid-latitudes. Besides seasonal and interannual variabilities, it is important to investigate the long-term tendencies in water vapor quantitatively. All climate models predict that there is an increase the stratospheric water vapor trends (Gettelman, 2010). Multi-year observations over mid-latitudes show that there is a positive trend in water vapor data considered for the years 1954–2000 (Rosenlof et al., 2001) and 1980–2010 (Hurst et al., 2011). However, there are only a few studies, which showed altitude dependence with positive trends in upper stratosphere and negative trends in lower stratosphere for the water vapor data extending back to 1986 (Scherer et al., 2008, Hegglin et al., 2014). Over the tropics, stratospheric water vapor for the years shows no significant long-term trend just above the tropopause (82 hPa) (Dessler et al., 2014).

In this paper, long-term trends and variabilities of upper tropospheric and lower stratospheric (UTLS) water vapor mixing ratio (WVMR) over the tropics (30°S–30°N) obtained by the Microwave Limb Sounder (MLS) instrument on board Aura Earth Observing System satellite (EOS) for the period October 2004–September 2015 are studied using multivariate regression analysis.

Section snippets

Water vapor mixing ratio (WVMR) data

The MLS instrument on board Aura EOS satellite was launched on 15 July 2004. It uses a sun-synchronous orbit at an altitude of 705 km and with 98° inclination. The standard water vapor product of v4.2 is taken from the 190 GHz channel. The horizontal grid is every 1.5° or ~165 km along the orbit track. The version 4.2 (v4.2) stratospheric water vapor data used in this study are the update to the v3.3 and v2.2 versions, which are validated and described by Hurst et al. (2014) and Lambert et al.,

WVMR trend

Fig. 4a shows the height profile of trend shown by the WVMR for the pressure levels 178-14.7 hPa obtained from the regression analysis described in Section 2. There is not much difference in the trends with and without the SC term. It is found that there is a decreasing trend of 0.02–0.1 ppmv/year in the WVMR below 100 hPa while the trend is positive (0.02–0.035±0.005 ppmv/year) above 100 hPa. There is no significant trend at 121 hPa. The maximum positive trend of 0.04±0.008 ppmv/year is observed at 68

Discussion and conclusion

In this study, long-term variabilities and tendencies in tropical UTLS WVMR and temperature obtained from the MLS instrument on board EOS satellite are studied using multivariate regression analysis. There is a decreasing trend of 0.02–0.1 ppmv/year in WVMR below 100 hPa while the trend is positive above 100 hPa with the maximum of 0.04 ppmv/year at 68 hPa. Almost all climate models predict that stratospheric WVMR will continue to increase in future (Gettelman et al., 2009). Oltmans and Hofmann

Summary

In this study, long-term variabilities and tendencies of tropical UTLS WVMR and temperature have been reported and discussed considering most of the important parameters, which could influence the WVMR variations. The WVMR shows an increasing (decreasing) trend in WVMR (temperature) above 100 hPa (56 hPa). There is no significant trend in WVMR and temperature near tropopause (100 hPa). The responses of WVMR and temperature to different parameters are discussed. The interannual variability of

Acknowledgments

The MLS team (NASA JPL) is gratefully acknowledged for their data. ERA-interim data used this study were provided by ECMWF and downloaded from their servers. The QBO winds, solar flux and MEI are downloaded from the websites http://www.geo.fu-berlin.de/met/ag/strat/produkte/qbo, ftp://ftp.geolab.nrcan.gc.ca/data/; http://www.esrl.noaa.gov/psd/enso/mei/table.html. The authors thank the Editor and the two Reviewers for their comments and suggestions, which greatly helped them to improve the

References (65)

  • V. Panwar et al.

    Some features of water vapor mixing ratio in tropical troposphere and lower stratosphere: role of convection

    Atmos. Res.

    (2012)
  • M.P. Baldwin et al.

    The Quasi-Biennial oscillation

    Rev. Geophys.

    (2001)
  • G. Beig et al.

    In search of greenhouse signal in the equatorial middle atmosphere

    Geophys. Res. Lett.

    (2001)
  • G. Beig et al.

    Solar response in temperature over the equatorial middle atmosphere

    J. Atmos. Sol. Terr. Phys.

    (2009)
  • A.W. Brewer

    Evidence for a world circulation provided by measurements of helium and water vapour distribution in the stratosphere

    Quart. J. R. Meteor. Soc.

    (1949)
  • J.M. Castanheira et al.

    Relationship between Brewer-Dobbson circulaton, double tropopauses, ozone and stratospheric water vapour

    Atmos. Chem. Phys.

    (2012)
  • A.E. Dessler et al.

    Stratospheric water vapor feedback

    PNAS

    (2013)
  • A.E. Dessler et al.

    Variations of stratospheric water vapor over the past three decades

    J. Geophys. Res. Atmos.

    (2014)
  • S. Dhomse et al.

    The relationship between tropospheric wave forcing and tropical lower stratospheric water vapor

    Atmos. Chem. Phys.

    (2008)
  • K. Drdla et al.

    Temperature thresholds for chlorine activation and ozone loss in the polar stratosphere

    Ann. Geophys.

    (2012)
  • S.J. Evans et al.

    Trends in stratospheric humidity and the sensitivity of ozone to these trends

    J. Geophys. Res.

    (1998)
  • S. Fadnavis et al.

    Seasonal variation of trend in temperature and ozone over the tropical stratosphere in the Northern Hemisphere

    J. Atmos. Sol. Terr. Phys.

    (2006)
  • P.M.F. de Forster et al.

    Stratospheric water vapour changes as a possible contributor to observed stratospheric cooling

    Geophys. Res. Lett.

    (1999)
  • P. Foster et al.

    Radiative forcing and temperature trends from stratospheric ozone changes

    J. Geophys. Res.

    (1997)
  • P.M. Foster et al.

    Assessing the climate impact of trends in stratospheric water vapour

    Geophys. Res. Lett.

    (2002)
  • S. Fueglistaler et al.

    Stratospheric water vapor predicted from the Lagrangian temperature history of air entering the stratosphere in the tropics

    J. Geophys. Res.

    (2005)
  • S. Fueglistaler et al.

    The tropical tropopause layer

    Rev. Geophys.

    (2009)
  • K.S. Gage et al.

    Solar variability and the secular variation in the tropical tropopause

    Geophys. Res. Lett.

    (1981)
  • A. Gettelman

    Multimodel assessment of the upper troposphere and lower stratosphere: tropics and global trends

    J. Geophys. Res.

    (2010)
  • A. Gettelman et al.

    The tropical tropopause layer 1960–2100

    Atmos. Chem. Phys.

    (2009)
  • M.A. Giorgetta et al.

    Potential role of the quasi-biennial oscillation in the stratosphere-troposphere exchange as found in water vapor in general circulation model experiments

    J. Geophys. Res.

    (1999)
  • S.A. Glantz et al.

    Primer of Applied Regression and Analysis of Variance

    (1990)
  • M.I. Hegglin et al.

    Vertical structure of stratospheric water vapour trends derived from merged satellite data

    Nat. Geosci.

    (2014)
  • J.R. Holton et al.

    Horizontal transport and the dehydration of the stratosphere

    Geophys. Res. Lett.

    (2001)
  • J.R. Holton et al.

    Stratosphere-troposphere exchange

    Rev. Geophys.

    (1995)
  • L.L. Hood

    Thermal response of the tropical tropopause regions to solar ultraviolet radiations

    Geophys. Res. Lett.

    (2003)
  • D.F. Hurst et al.

    Stratospheric water vapor trends over Boulder, Colorado: analysis of the 30 year Boulder record

    J. Geophys. Res.

    (2011)
  • D.F. Hurst et al.

    Validation of Aura Microwave Limb Sounder stratospheric water vapor measurements by the NOAA frost point hygrometer

    J. Geophys. Res.

    (2014)
  • R.L. Jones et al.

    Observations of CH4 and N2O by the NIMBUS 7 SAMS: a comparison with in situ data and two-dimensional numerical model calculations

    J. Geophys. Res.

    (1984)
  • Y. Kawatani et al.

    Interannual variations of stratospheric water vapor in MLS observations and climate model simulations

    J. Atmos. Sci.

    (2014)
  • K. Kodera et al.

    Dynamical response to the solar cycle

    J. Geophys. Res.

    (2002)
  • A. Kunz et al.

    Extending water vapor trend observations over Boulder into the tropopause region: trend uncertainties and resulting radiative forcing

    J. Geophys. Res.

    (2013)
  • Cited by (5)

    • Dry air valley in the upper troposphere over the eastern Mediterranean-western Tibetan Plateau

      2023, Atmospheric Research
      Citation Excerpt :

      A number of studies have shown that the Tibetan Plateau (TP) and tropics are key regions for the transport of water vapor from the upper troposphere to the stratosphere (Gettelman et al., 2004; Schoeberl et al., 2013; Qie et al., 2022), and stratospheric water vapor changes can significantly affect the tropospheric climate through radiation and chemical processes (Solomon et al., 2010). Hence, previous studies mainly focused on the upper tropospheric water vapor over the TP and tropics (e.g., Panwar et al., 2012; Sridharan and Sandhya, 2016; Xu et al., 2021, 2022). Tian et al. (2011) found that there is a dry air valley (DAV) in the upper troposphere over the western Tibetan Plateau during the period of May to September, which extend westward to the eastern Mediterranean in summer months.

    • Seasonal, interannual and long-term variabilities and tendencies of water vapour in the upper stratosphere and mesospheric region over tropics (30°N-30°S)

      2018, Journal of Atmospheric and Solar-Terrestrial Physics
      Citation Excerpt :

      Recently, using 14 years (1991–2005) time series of mesospheric water vapour measurements obtained from HALOE for 45°S-45°N, Remsberg (2010) found that the annual average water vapour mixing ratio decreases from 6.5 ppmv at 0.2 hPa to 3.2 ppmv at 0.01 hPa, in correlation with Lyman-α flux. There are a few studies (Russel et al., 1993; Mote et al., 1996; Harris et al., 1996; Kley et al., 2000; Jones et al., 2009; Nedoluha et al., 2009; Li et al., 2011; Hegglin et al., 2013; Sridharan and Sandhya, 2016) on seasonal and long-term water vapour variations in the stratosphere and mesosphere using satellite data sets. The present study uses the long-term (October 2004–September 2015) Tropical water vapour measurements for the height region 31–102 km obtained from the Microwave Limb Sounder (MLS) data sets onboard Aura satellites to study the seasonal and interannual variabilities.

    View full text