Paper_35yr_Jan31

advertisement
1
2
3
El Niño-Southern Oscillation in Tropical Column Ozone and A
4
3.5-year signal in Mid-Latitude Column Ozone
5
6
Jingqian Wang,1* Steven Pawson,2 Baijun Tian,3 Maochang Liang,4 Run-Lie Shia,5
7
Yuk L. Yung,5 and Xun Jiang1
8
9
10
11
12
13
1 Department of Earth and Atmospheric Sciences, University of Houston, TX 77204,
14
USA.
15
2 Global Modeling and Assimilation Office, NASA GSFC, Code 610.1, Greenbelt, MD
16
20771, USA.
17
18
19
3 Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr.,
Pasadena, CA 91109, USA.
4 Research Center for Environmental Changes, Academia Sinica, Taipei, Taiwan.
20
21
22
5 Division of Geological and Planetary Sciences, California Institute of Technology,
1200 East California Boulevard, Pasadena, CA 91125, USA.
23
* To whom all correspondence should be addressed. E-mail: jwang3@mail.uh.edu
24
25
To be submitted to JGR, Jan 27, 2010
26
-1-
27
Abstract
28
[1] Impact of El Niño-Southern Oscillation (ENSO) on the tropical column ozone,
29
tropical tropopause pressure, and the 3.5-year ozone signal [first identified by Jiang et al.
30
2008] in the mid-latitude column ozone are examined by Version 1 of the Goddard Earth
31
Observation System Chemistry-Climate Model (GEOS CCM) [Pawson et al., 2008].
32
Observed monthly-mean sea surface temperature and sea ice between 1951 and 2004
33
were used as boundary conditions for the model. ENSO signal is the dominant mode in
34
the tropical column ozone variability in the GEOS CCM, capturing 65.8% of the total
35
variance. The spatial pattern of this mode is similar to that in Total Ozone Mapping
36
Spectrometer (TOMS) observations between Nov 1978 and Apr 1993. However, there
37
are some discrepancies in ozone spatial pattern in the southern hemisphere between the
38
GEOS CCM and TOMS. There is also a clear ENSO signal in the tropical tropopause
39
pressure in the GEOS CCM, which may be the cause of the ENSO signal in the column
40
ozone in the GEOS CCM. The regression coefficient between the model O3 and the
41
model tropopause pressure is 0.78 DU/hPa.
42
43
[2] GEOS CCM is also used in this paper to investigate possible mechanism for the 3.5-
44
year signal in the mid-latitude column ozone. Since the 3.5-year signal also appears in the
45
GEOS CCM column ozone as those in the ozonesondes, it suggests that a model with
46
realistic ENSO can reproduce the 3.5-year signal. Hence, it is likely that the 3.5-year
47
signal is from the ENSO signal.
48
49
-2-
50
1. Introduction
51
[3] The influence of El Niño-Southern Oscillation (ENSO) on the sea surface
52
temperature (SST), surface pressure, winds, and convection has been well known [see,
53
e.g., Trenberth and Shea, 1987; Trenberth, 1997]. ENSO can also affect the column
54
ozone abundance. Bojkov [1987] first found low column ozone anomalies in the middle
55
and high latitudes within a few months after El Niño episodes. Shiotani [1992] studied
56
the asymmetric ENSO signal in the equatorial column ozone by removing the zonal
57
means. Kayano [1997] carried out an empirical orthogonal function (EOF) study of the
58
total ozone between 70º N and 70º S and found that the first two modes are related to
59
ENSO. Zerefos [1992] found a correlation between the total column ozone and the ENSO
60
in the tropical troposphere. The decreases in total ozone in middle and sometimes polar
61
latitude follow the ENSO by several months. It has been proposed that the warming in the
62
tropics over the Pacific Ocean causes increases in the upper troposphere air temperature
63
and tropopause height and an upwelling in the lower stratosphere [Kalicharran, 1993].
64
More recently, Camp et al. [2003] used the column ozone observed by Total Ozone
65
Mapping Spectrometer (TOMS) and revealed that there is an ENSO signal in the fourth
66
mode of the column ozone in the tropical region. During the El Niño periods, there are
67
negative column ozone anomalies in the eastern tropical Pacific and positive column
68
ozone anomalies in the western Pacific. The influence of ENSO on column ozone can
69
mostly be explained by the variation of tropopause height driven by changes of tropical
70
deep convection [Schubert and Munteanu, 1988; Shiotani, 1992; Hasebe, 1993]. Such a
71
causal relationship between column O3 and the tropopause height is also evident in the
72
Madden-Julian Oscillation, as recently pointed out by Tian et al. [2007]. However, a
-3-
73
quantitative demonstration of the impact of ENSO on column O3 and tropopause height
74
has not been carried out using model simulations.
75
76
[4] In this paper, we will focus on the ENSO signal in the tropopause pressure and
77
column ozone simulated by Version 1 of the Goddard Earth Observation System
78
Chemistry-Climate Model (CCM) [Pawson et al., 2008], and try to investigate the
79
mechanism for the mid-latitude column ozone 3.5-year signal. The 3.5-year signal is first
80
found in Arosa ozonesonde. The same signal is also found in the first leading mode of
81
column ozone in the northern hemisphere [Jiang et al., 2008]. However, mechanism for
82
the 3.5-year signal in ozone is not known, which will be explored in this paper.
83
84
2. GEOS-4 Chemistry-climate Model and Data
85
[5] The GEOS CCM (Version 1) [Stolarski et al., 2006; Pawson et al., 2008] combines
86
the GEOS-4 general circulation model [Bloom et al., 2005] with a stratospheric chemistry
87
model [Douglass et al., 2003; Stolarski et al., 2006]. Interaction between atmospheric
88
composition and the circulation is through use of predicted O3, H2O, CO2, CH4, N2O,
89
CFC-11, and CFC-12 in the radiative heating computation. The model run analyzed in
90
this study, which is the run P1 of Pawson et al. [2008], spanned 54 years (1951-2004) at
91
2º 2.5º (latitude by longitude) with 55 layers between the surface and about 80km. At
92
the lower boundary, monthly mean sea surface temperature and sea ice are prescribed
93
from observations [Rayner et al., 2003], along with time-dependent emissions of
94
greenhouse gases and chloroflorocarbons, as described in Eyring et al. [2006]. Neither
95
the solar cycle nor volcanic aerosol variations were included in the model and it does not
-4-
96
capture a QBO [see Horinouchi et al., 2003], so the signals of these sources in ozone
97
variability are all absent from the simulation.
98
99
[6] Isolation of the ENSO signal in the GEOS CCM will be performed using gridded
100
model output on a 2 latitude by 2.5 longitude grid. Two simulated fields will be
101
examined: the tropopause pressure and the column ozone. The tropopause pressure, PT,
102
was determined from the simulations using the lapse-rate definition of tropopause.
103
Column ozone, ΩS, is the integrated profile of the total amount of ozone in atmosphere. A
104
corresponding analysis of the observed SSTs between 1951 and 2004 will also be
105
performed.
106
107
[7] Ozonesonde data, which have relatively long data record, will be investigated in this
108
paper. (Data can be downloaded from http://www.woudc.org/data_e.html). Twenty-one
109
stations with long records will be used in this paper. These ozonesonde data will be
110
compared with model simulation from GEOS CCM.
111
112
3. Methods
113
[8] The following filtering and analysis techniques were applied to three time series of
114
gridded, monthly mean data. No spatial smoothing was performed. First, the mean annual
115
cycle was computed by independently evaluating the 54-year mean value for each month.
116
This mean annual cycle and the linear trend were subtracted from the original time series
117
to create the detrended monthly anomalies for each month. Second, a spectral filter was
118
applied to the monthly anomalies, to isolate the interannual variability (IAV) from higher
-5-
119
frequency oscillations. The filter was constructed as the convolution of a step function
120
with a Hanning window and chosen to obtain a full signal from periods above 15 months
121
and no signal from periods below 12.5 months.
122
123
[9] Finally, principal component analysis (PCA) [Preisendorfer, 1988] was performed on
124
the filtered tropical (25 ºS to 25 ºN) anomalies of SST, ΩS, and PT. PCA provides a
125
decomposition of a multivariate dataset into orthogonal spatial patterns, known as
126
empirical orthogonal functions (EOF), with associated time-dependent amplitude, known
127
as principal component time series (PC). The EOFs are the eigenfunctions of the
128
covariance matrix of the dataset sorted by the decreasing values of associated
129
eigenvalues. Since these eigenvalues represent the variance captured by each EOF, PCA
130
guarantees that the leading EOFs capture most of the variance of the dataset.
131
132
[10] For the observation data (ozonesonde data), We first remove the seasonal cycle
133
from the data then remove the trend. Power spectral analysis is applied to deseasonalized
134
and detrended ozonesondes. Spectral analysis will also be applied to the deseasonalized
135
model O3 simulations. To obtain the statistical significance of signals in a power
136
spectrum, we will compare the amplitude of a spectral peak to the red noise spectrum
137
[Gilman et al., 1963]. The 15%, 10%, and 5% significance levels for the power spectrum
138
are found using F-statistics to compare the spectrum to the red noise spectrum.
139
140
4. ENSO Signals and 3.5-year Column Ozone Signal
141
4.1 ENSO Signals
-6-
142
[11] Principal component analysis (PCA) [Preisendorfer, 1988] was performed on the
143
filtered tropical (25 ºS to 25 ºN) anomalies of SST, ΩS, and PT. The first modes of
144
variability determined by the PCA capture 46.9%, 33.9%, and 63.8% of the total variance
145
in SST, PT, and ΩS, respectively (see Table 1). Evidence is presented that the first modes
146
of variability captured by the PCA are related to ENSO. The Southern Oscillation Index
147
(the pressure difference between Tahiti and Darwin) is used to establish this
148
correspondence. Visual comparison of time series of the SOI and the amplitude of the
149
first mode (PC1) of the SST (Fig 1) reveals a good correspondence, with in-phase
150
variations. Note that negative/positive SOI values denote the El Niño/La Niña events.
151
The correlation between the two time series is 0.88 and is significant at the 0.1% level.
152
The significance level was generated by a Monte Carlo method [Press et al., 1992]: a
153
small numerical value denotes high statistical confidence. Time series of PC1 for P T and
154
ΩS reveal a similar low-frequency variability as for SOI, with some higher-frequency
155
variations superimposed. This leads to somewhat lower correlations (0.53 and 0.65)
156
between these time series and the SOI, but these remain highly significant (see Table 1).
157
The correlation between PC1s of PT and ΩS is 0.82. Power spectra of the PC1s of SST, PT
158
and ΩS (Fig. 2) all reveal strong spectral peaks near 3-7 years, which are typical of ENSO
159
variations. The higher frequency variations evident in PT and ΩS also lead to prominent
160
peaks near 17 months, which might be a beat frequency between annual cycle and ENSO
161
signal.
162
163
[12] Figure 3 shows the spatial patterns of the first modes of the PCA analysis for the
164
three fields, which describe their ENSO-related variability. For SST, the spatial pattern
-7-
165
(Fig. 3a) is similar to the first mode of ENSO-related variability described in Trenberth et
166
al. [2002], which corresponds to the SST anomalies in the Niño-3.4 region. Negative
167
amplitudes of this mode correspond to the El Niño events, with warm SST anomalies in
168
the central and eastern Pacific and cold anomalies in the western Pacific. During El Niño
169
periods, the convection will move eastward to the central Pacific and impact the
170
longitudinal structure of the tropopause [e.g., Gage and Reid, 1987]. The spatial structure
171
of the first mode of PT (Fig. 3b) is consistent with this behavior. For the negative
172
amplitudes of this mode, corresponding to El Niño events, there are large negative
173
tropopause pressure anomalies (higher tropopause height) in the central and eastern
174
Pacific and weak positive tropopause pressure anomalies (lower tropopause heights) in
175
the equatorial western Pacific. Please note the tropopause pressure anomalies are
176
relatively small over the equatorial regions. On the other hand, the large tropopause
177
pressure anomalies are found over the subtropics, which is the Rossby-wave response of
178
the equatorial ENSO convection anomaly [Trenberth et al., 1998]. In the positive
179
amplitude of this mode, that is, La Niña events, there are weak negative tropopause
180
pressure anomalies (higher tropopause heights) in the equatorial western Pacific and large
181
positive tropopause pressure anomalies (lower tropopause heights) in the central and
182
eastern Pacific. The present results are in general agreement with previous studies [e.g.,
183
Kiladis et al., 2001; Gettelman et al. 2001] on the tropopause pressure and height
184
variability associated with the ENSO based on reanalysis data. This indicates that the
185
vertical movement of the tropopause associated with the ENSO is reasonably captured by
186
the GEOS CCM.
187
-8-
188
[13] One important aspect of the spatial pattern of the PT anomaly is that it has the same
189
sign everywhere except for a confined region centered over the Equator to the North of
190
Indonesia (Fig. 3b). This contrasts with the existence of positive and negative anomalies
191
in the SST pattern. This means that almost the entire tropical tropopause is displaced
192
upwards (to lower pressure) in El Niño events, with a discernable signal in the zonal
193
mean. A similar result is true for the ΩS pattern (Fig. 3c), in that substantial zonal
194
anomalies are superimposed on a monotonic signal. Note that the ENSO signal is the
195
dominant pattern of variability in ΩS because irradiance changes associated with the solar
196
cycle were not imposed in the simulation and because this version of GEOS CCM does
197
not simulate a QBO. The pattern is very similar to the ENSO spatial pattern in the TOMS
198
column ozone data in Camp et al. [2003], although there are some discrepancies in the
199
southern hemisphere. The spatial pattern is also consistent with the second mode (mature
200
stages of the El Niño) for column ozone from Kayano et al. [1997]. There is a double-
201
ridged symmetric structure with ridges at 12º N and 15º S. The value is relatively low in
202
the western Pacific and relatively high in the eastern Pacific.
203
204
[14] A broadly consistent picture arises of the leading-order anomalies, with (in El Niño
205
events) an eastward migration of warm SST anomalies that enhance convection over the
206
eastern Pacific where the tropopause rises and ΩS decreases. This decrease in column
207
ozone is consistent with the reduction in depth of the stratophere. It is also consistent
208
with a faster vertical flow through the ozone source region in the tropical stratosphere,
209
which tends to reduce the ozone column.
210
-9-
211
[15] We also regress the first mode of the column ozone on the first mode of the
212
tropopause pressure. The mean value for the first mode of the column ozone in the Niño-
213
3.4 region is plotted against the mean value for the first mode of the tropopause pressure
214
in Fig. 4a. These two time series correlate well. The correlation coefficient is 0.82. The
215
scatter plot for the two time series is shown in Fig. 4b. The regression coefficient
216
between the O3 and tropopause pressure is 0.71 DU/hPa. Assuming that the tropopause
217
height is 10.5 km and scale height is 7 km, the regression coefficient is about -15.5
218
DU/km, which is consistent with the previous results varying from -10 to -25 DU/km
219
[Meetham, 1937; Reed, 1950; Hoinka et al., 1996; Steinbrecht et al., 1998].
220
221
4.2 3.5-year Ozone signal
222
[16] There were 21 observation stations that have long time record, but only eight
223
stations have significant 3.5-year signal. These eight stations are Potsdam, Belsk, Hradec,
224
Hohenpeissenberg, Arosa, Toronto, Sappor, Nashiville (see Table 2). We layout the
225
observation stations and their positions, and found the signal has no preference for the
226
spatial distributions. Power spectra and statistical significances of the peaks are shown in
227
Figure 5 for the eight stations. Dotted lines are the mean red-noise spectra. Dash-dot lines
228
and dashed lines correspond to 10% and 5% significance levels, respectively. In Fig. 5a,
229
there are about five significant spectral peaks (10% significance level) in the low-
230
frequency regions: residual annual cycle, QBO-Annual beat, QBO, 3.5-year signal, and
231
decadal oscillation. Since the mechanisms for the QBO-Annual beat, QBO, and decadal
232
oscillation are well known, we will only focus on the 3.5-year signal in this paper.
233
Among eight stations, there are significant peaks between 3-4 years. The power spectra
- 10 -
234
of Fig. 5a-Potsdam, d-Hohenpeissenberg, and g-Sappor are within 10% to 5%
235
significance levels. The 3.5-year signals in the other five stations, Fig. 5b-Belsk, c-
236
Hradec, e- Arosa, f-Toronto, h-Nashiville, are within 5% significance level.
237
238
[17] In order to reveal the mechanism for the 3.5-year cycle and explain the results from
239
the above, we investigated the signal in GEOS CCM. GEOS CCM is forced by realistic
240
sea surface temperature and sea ice. The ENSO signal is well captured at the model
241
surface. There is no QBO and solar cycle in the model. We applied the same spectral
242
analysis to the model output. Results for model column ozone are shown in Figure 6.
243
There are significant 3.5-year signals in all stations. The significance levels are within
244
10%-5% for column ozone at Potsdam, Belsk, Hradec, Hohenpeissenberg, Arosa, and
245
Sapporo (Figs. 6a-6e and Fig. 6g). At Toronto and Nashville (Fig. 6f and Fig. 6h), the
246
significance levels are within 5% for column ozone. It is obvious that the 3.5-year signal
247
could also be found in the model column ozone and is significant. Since the model only
248
has realistic interannual variability from ENSO and no other interannual variability is
249
available, it suggests that the 3.5-year signal in ozone might origin from ENSO signal
250
from surface.
251
252
5. Conclusion
253
[18] In this paper, we have studied the ENSO signal in the tropical column ozone and
254
tropical tropopause pressure in the GEOS CCM from Jan 1951 to Dec 2004. The first
255
modes in the model column ozone and tropopause pressure capture 63.8% and 33.9% of
256
the total variances, respectively. They are related to ENSO. PC1 for the first modes in the
- 11 -
257
model column ozone and tropopause pressure correlate well with the SOI index and
258
leading PC time series for the SST. The spatial pattern of the first mode for the model
259
column ozone is similar to the ENSO signal in the TOMS column ozone, although there
260
are some discrepancies for the ozone in the southern hemisphere. The regression
261
coefficient between the model O3 and the model tropopause pressure is consistent with
262
observations. It can be improved in the future with the inclusion of tropospheric
263
chemistry in the chemical module of the GEOS CCM.
264
265
[19] In the middle latitude, we apply spectral analysis to the observed monthly mean
266
ozonesonde data and find the 3.5-year signal in eight stations. The 3.5-year signal is also
267
found in the GEOS CCM model ozone. Since GEOS-CCM only has realistic ENSO and
268
there is no QBO and solar cycle signals in the model, it suggests that the 3.5-year signal
269
in column ozone is related to the ENSO signal. The linkage between 3.5-year ozone
270
signal and surface ENSO needs to be investigated in more details in the future.
271
272
[20] Acknowledgments. We specially acknowledge Dr. Rappenglueck and Dr. Lefer,
273
who gave great commends and help in this research.
274
- 12 -
275
276
References
277
Bloom, S., et al. (2005), The Goddard Earth Observation System Data Assimilation
278
System, GEOS DAS Version 4.0.3: Documentation and Validation, NASA TM-
279
2005-104606, 26, 166.
280
281
Bojkov, R.D. (1987), The 1983 and 1985 anomalies in ozone distribution in perspective,
Mon. Weather Rev., 115, 2187-2201.
282
Camp, C.D., M.S. Roulston, and Y.L. Yung (2003), Temporal and Spatial Patterns of the
283
Interannual Variability of Total Ozone in the Tropics, J. Geophys. Research-
284
Atmospheres, 108, art. no. 4643.
285
Douglass, A. R., et al. (2003), Evaluation of transport in the lower tropical stratosphere in
286
a global chemistry and transport model, J. Geophys. Research-Atmospheres, 108,
287
NO. D9, 4259, doi:10.1029/2002JD002696.
288
Eyring, V., et al. (2006), Assessment of temperature, trace species, and ozone in
289
chemistry-climate model simulations of the recent past, J. Geophys. Res., 111,
290
D22308, doi:10.1029/2006JD007327.
291
Fusco, A.C., and M.L. Salby (1999), Interannual variations of total ozone and their
292
relationship to variations of planetary wave activity, Journal of Climate, 12, 1619-
293
1629.
294
Gage, K.S., and G.C.Reid (1987), Longitudinal variations in tropical tropopause
295
properties in relation to tropical convection and El-Niño-Southern Oscillation, J.
296
Geophys. Res., 92, 14,197-14,203.
- 13 -
297
Gettelman, A., W.J.Randel, S. Massie, F. Wu, W.G. Read, and J.M. Russell III (2001),
298
El-Niño as a natural experiment for studying the tropical tropopause region, J.
299
Clim., 14, 3375-3392.
300
301
302
303
304
305
Gilman D.L., F. J. Fuglister, and J.M. Mitchell Jr. (1963), On the Power Spectrum of
“Red Noise”, J. Atmos. Sci., 20, 182-184.
Hasebe, F. (1993), Dynamical response of the tropical total ozone to sea surface
temperature changes, J. Atmos. Sci., 50, 345-356.
Hoinka, K.P., H. Claude, and U. Kohler (1996), On the correlation between tropopause
pressure and ozone above central Europe, Geophys. Res. Lett., 23, 1753-1756.
306
Horinouchi, T., S. Pawson, K. Shibata, U. Langematz, E. Manzini, M.A. Giorgetta, F.
307
Sassi, R.J. Wilson, K.P. Hamilton, J. de Grandpré, A.A. Scaife (2003), Tropical
308
Cumulus Convection and Upward Propagating Waves in Middle Atmospheric
309
GCMs, J. Atmospheric Sci., 60, 2765-2782,.
310
Huang, N.E., et al. (1998), The empirical mode decomposition and the Hilbert spectrum
311
for nonlinear and non-stationary time series analysis, The Royal Society, 454, 904-
312
993.
313
Jiang, X., C. D. Camp, R. Shia, D. Noone, C. Walker, and Y. L. Yung (2004), Quasi-
314
biennial oscillation and quasi-biennial oscillation-annual beat in the tropical total
315
column ozone: A two-dimensional model simulation, J. Geophys Res, 109,
316
doi:10.1029/2003JD004377.
317
Jiang, X., S. Pawson, C. D. Camp, E. Nielsen, R. Shia, T. Liao, K. Jeev, V. Limpasuvan,
318
and Y. L. Yung (2008), Interannual variability and trends in extratropical ozone.
319
Part I: Northern hemisphere, J. Atmos. Sci., 65, 3013-3029.
- 14 -
320
Jiang, X., S. Pawson, C. D. Camp, E. Nielsen, R. Shia, T. Liao, K. Jeev, V. Limpasuvan,
321
and Y. L. Yung (2008), Interannual variability and trends in extratropical ozone.
322
Part II: Southern hemisphere, J. Atmos. Sci.,65, 3030-3041.
323
324
Kalicharran, S., R. D. Diab, and F. Sokolic (1993), Trends in total ozone over Southern
African stations between 1979 and 1991, Geophys. Res. Lett, 20, 2877-2880.
325
Kayano, M.T. (1997), Principal modes of the total ozone on the Southern Oscillation
326
timescale and related temperature variations, J. Geophys. Res., 102 (D22), 25797-
327
25806.
328
Kiladis GN, Straub KH, Reid GC, et al. (2001), Aspects of interannual and intraseasonal
329
variability of the tropopause and lower stratosphere, Quart. J. Roy. Meteor. Soc.,
330
127, 1961-1983
331
Ma, J., D. W. Waugh_ A. R. Douglass, S. R. Kawa, P. A. Newman, S. Pawson, R.
332
Stolarski and S. J. Lin (2004), Interannual variability of stratospheric trace gases:
333
Role of extratropical wave driving, Quart. J. Roy. Meteor. Soc., 128, 1-999.
334
Meetham, A. R. (1937), The correlation of the amount of ozone with other characteristics
335
of the atmosphere, Quart. J. Roy. Meteor. Soc., 63, 289-307.
336
Pawson, S., R. S. Stolarski, A. R. Douglass, P. A. Newman, J. E. Nielsen, S. M. Frith,
337
and M. L. Gupta (2008), Goddard Earth Observing System chemistry-climate
338
model simulations of stratospheric ozone-temperature coupling between 1950 and
339
2005, J. Geophys. Res., 113, D12103, doi:10.1029/2007JD009511.
340
341
Preisendorfer, R. W. (1988), Principal component analysis in meteorology and
oceanography, Development in Atmospheric Science, 17. 24 cm. 425 p.
- 15 -
342
Press, W., S. Teukolsky, W. Vetterling, and B. Flannery (1992) , Numerical Recipes in
343
Fortran 77: The Art of Scientific Computing, 933 pp., Cambridge Univ. Press,
344
New York.
345
346
Randel., W.J., and J.B.Cobb (1994), Coherent variations of monthly mean total ozone
and lower stratospheric temperatures. J. Geophys. Res., 99, 5433-5447.
347
Randel, W.J., F. Wu, and R. Stolarski, (2002), Changes in column ozone correlated with
348
the stratospheric EP flux, J.ournal of the Meteorological Society of Japan, 80,
349
849-862.
350
Rayner, N. A., et al. (2003), Global analyses of sea surface temperature, sea ice, and
351
night marine air temperature since the late nineteenth century, J. Geophys.
352
Research-Atmospheres, 108.
353
354
Reed, R. J. (1950), The role of vertical motions in ozone-weather relationship, J. Meteor.,
7, 263-267.
355
Rex, M., R.J. Salawitch, P. von der Gathen, N.R.P. Harris, M.P. Chipperfield, and B.
356
Naujokat (2004), Arctic ozone loss and climate change, Geophys. Res. Lett., 31,
357
L04116, doi: 10.1029/2003GL018844.
358
Reynolds R. W., N. A. Rayner, T. M. Smith, D. C. Stokes and W. Wang (2002), An
359
Improved In Situ and Satellite SST Analysis for Climate, J. Climate, 15, 1609-
360
1625.
361
Richman, M.B. (1986), Rotation of principal components. J. Climatology, 6, 293-335.
362
Schubert, S.D., and M.J. Munteanu (1988), An analysis of tropopause pressure and total
363
ozone correlations, Mon. Weather Rev., 116, 569-582.
- 16 -
364
Scott, R.K., and P.H. Haynes (1998), Internal interannual variability of the extratropical
365
stratospheric circulation: The low-latitude flywheel, Q.uart J. Roy. Meteor. Soc.,
366
124, 2149-2173.
367
Shiotani, M. (1992), Annual, quasi-biennial, and El Niño-Southern Oscillation (ENSO)
368
timescale variations in equatorial total ozone, J. Geophys. Res., 97 (D7), 7625-
369
7633.
370
Steinbrecht, W., H. Claude, U. Kohler, and K. P. Hoinka (1998), Correlations between
371
tropopause height and total ozone: Implications for long-term changes, J.
372
Geophys. Res., 103 (D15), 19183-19192.
373
374
Stolarski, R.S., A.R. Douglass, S.E. Steenrod, S. Pawson (2006), Trends in Stratospheric
1028-1041.
375
Tian, B., et al. (2007), Intraseasonal variations of the tropical total ozone and their
376
connection to the Madden-Julian Oscillation, Geophys. Res. Lett., 34, L08704,
377
doi:10.1029/2007GL029451.
378
Trenberth, K.E. (1998), Progress during TOGA in understanding and modeling global
379
teleconnections associated with tropical sea surface temperatures. J. Geophys.
380
Res., 103, 14291-14324.
381
382
Trenberth, K.E., and D.P. Stepaniak (2001), Indices of El Nino evolution, J . Climate, 14,
1697-1701.
383
Trenberth, K.E., J.M. Caron, D.P. Stepaniak, and S. Worley (2002), Evolution of El
384
Niño-Southern Oscillation and global atmospheric surface temperatures, J.
385
Geophys. Res., 107 (D8), doi: 10.1029/2000JD000298.
- 17 -
386
387
Trenberth, K.E. and D. J. Shea (1987), On the evolustion of the Southern Osillation
envent. Mon. Weather Rev., 115, 3078-3096.
388
Zerefos, C. S., A. F. Bais, I. C. Ziomas, and R. D. Bojkov (1992), On the Relative
389
importance of Quasi-Biennal Osicillation and ELNino/Southern Oscillation in the
390
Revised Dobson total Ozone Records, J. Geophs. Res., 97, 10135-10144.
391
392
393
- 18 -
394
Table 1: Variances, spectral peaks, and correlations (Lag = 0) for the first modes of the
395
sea surface temperature, tropopause pressure, and column ozone. The numbers in
396
parentheses denote significance levels.
397
Mode 1
Variance
Spectral
Correlations (significance level) with:
of
captured
Peaks
SOI
SST (PC1)
PT (PC1)
SST
46.9%
4-5 years
0.88 (0.1%)


PT
33.9%
17 months
0.65 (0.1%)
0.72 (0.1%)

0.53 (0.1%)
0.57 (0.1%)
0.82
5 years
ΩS
63.8%
17 months
2-5 years
398
399
- 19 -
400
Table 2. The observation stations that have 3.5-year column ozone signal.
Stations
Latitude
Longitude
Potsdam
52.36
13.08
Belsk
51.84
20.79
Hradec
50.18
5.83
Hohenpeissenberg
47.8
11.02
Arosa
46.7
9.68
Toronto
43.78
-79.47
Sapporo
43.05
141.33
Nashville
36.25
86.57
401
402
- 20 -
403
404
Figure 1: PC1 time series for the first modes of sea surface temperature (black solid line),
405
tropopause pressure (blue dotted line), and column ozone (red dashed line). SOI index is
406
shown by green dash-dotted line.
407
- 21 -
408
409
Figure 2: (a) Power spectrum of PC1 for the first mode of SST from GEOS-4 CCM. (b)
410
Power spectrum of PC1 for the first mode of tropopause pressure from GEOS-4 CCM.
411
(c) Power spectrum of PC1 for the first mode of column ozone from GEOS-4 CCM.
412
- 22 -
413
414
415
416
417
418
419
Figure 3: (a) The spatial pattern of the first mode for sea surface temperature in the
tropics. Units are K. The first mode explains 46.9% of the total variance. (b) The spatial
pattern of the first mode for tropopause pressure from GEOS-4 CCM in the tropics. Units
are Pa. The first mode explains 33.9% of the total variance. (c) The spatial pattern of the
first mode for column ozone from GEOS-4 CCM in the tropics. Units are DU. The first
mode explains 65.8% of the total variance.
- 23 -
420
421
422
423
424
425
Figure 4: (a) First mode of column O3 averaged in Niño-3.4 region (solid line) and first
mode of tropopause pressure in Niño-3.4 region (dotted line). (b) Scatter plot of O3
against the tropopause pressure. Solid line is the linear fitting of the dots.
- 24 -
426
427
428
Figure 5: Power spectra of ozonesonde data: (a) Potsdam, (b) Belsk, (c) Hradec, (d)
Hohenpeissenberg, (e) Arosa, (f) Toronto, (g) Sapporo, (h) Nashville.
- 25 -
429
430
431
432
Figure 6: Power spectra of model column ozone data simulated by GEOS CCM sampled
in the same locations as those in Figure 5.
- 26 -
Download